Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Binary mergers in the centers of galaxies: synergy between stellar flybys and tidal fields

Mila Winter-Granic Instituto de Astrofísica, Pontificia Universidad Católica de Chile, Av. Vicuña Mackenna 4860, 782-0436 Macul, Santiago, Chile Cristobal Petrovich Instituto de Astrofísica, Pontificia Universidad Católica de Chile, Av. Vicuña Mackenna 4860, 782-0436 Macul, Santiago, Chile Millennium Institute of Astrophysics MAS, Nuncio Monseñor Sótero Sanz 100, Of. 104, 750-0000 Providencia, Santiago, Chile Department of Astronomy, Indiana University, Bloomington, IN 47405, USA Valentín Peña-Donaire Instituto de Astrofísica, Pontificia Universidad Católica de Chile, Av. Vicuña Mackenna 4860, 782-0436 Macul, Santiago, Chile Chris Hamilton Institute for Advanced Study, Einstein Drive, Princeton, NJ 08540, USA
Abstract

Galactic centers are very dynamically active environments, often harbouring a nuclear star cluster and supermassive black hole at their cores. Binaries in these environments are subject to strong tidal fields that can efficiently torque its orbit, exciting near unity eccentricities that ultimately lead to their merger. In turn, frequent close interactions with passing stars impulsively perturb the orbit of the binary, generally softening their orbit until their evaporation, potentially hindering the role of tides to drive these mergers. In this work, we study the evolution of compact object binaries in the galactic center and their merger rates, focusing for the first time on the combined effect of the cluster’s tidal field and flyby interactions. We find a significant synergy between both processes, where merger rates increase by a factor of 1030similar-toabsent1030\sim 10-30∼ 10 - 30 compared to models in which only flybys or tides are taken into account. This synergy is a consequence of the persistent tides-driven eccentricity excitation that is enhanced by the gradual diffusion of jzsubscript𝑗𝑧j_{z}italic_j start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT driven by flybys. The merger efficiency peaks when the diffusion rate is 10100similar-toabsent10100\sim 10-100∼ 10 - 100 slower than the tides-driven torquing. Added to this synergy, we also find that the gradual softening of the binary can lift the relativistic quenching of initially tight binaries, otherwise unable to reach extreme eccentricities, and thus expanding the available phase-space for mergers. Cumulatively, we conclude that despite the gradual softening of binaries due to flybys, these greatly enhance their merger rates in galaxy centers by promoting the tidal field-driven eccentricity excitation.

galactic center – compact objects – star clusters – supermassive black holes – stellar dynamics

1 Introduction

Most nearby galaxies are known to harbor very dense stellar clusters at their core (Neumayer et al., 2011; Turner et al., 2012; Georgiev & Böker, 2014). These nuclear star clusters (NSCs) tend to have masses of roughly 105108Msuperscript105superscript108subscript𝑀direct-product10^{5}-10^{8}M_{\odot}10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, a small effective radii of only a few parsecs, and usually have a supermassive black hole (SMBH) residing at their center. This is the case for our own galaxy, containing an NSC with a mass of 3×107Msimilar-toabsent3superscript107subscript𝑀direct-product\sim 3\times 10^{7}M_{\odot}∼ 3 × 10 start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT (Schödel et al., 2014, 2018) and the central SMBH SgrA with a mass of about 4×106M4superscript106subscript𝑀direct-product4\times 10^{6}M_{\odot}4 × 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT (Ghez et al., 2005; Gillessen et al., 2009). This makes the galactic center (GC) a very dense and dynamical environment, which makes it an ideal laboratory to test several astrophysical phenomena involving stellar dynamics, general relativity (GR), and gravitational waves (GWs).

In the last decade there has been an increasing interest in gravitational-wave astrophysics, due to GW detections from the LIGO and VIRGO interferometers. Compact object binary mergers are known to be one of the main GW emitting events (Abbott et al., 2016, 2017), therefore recent studies have focused on the dynamical mechanisms that lead to these mergers to better understand the sources of GWs. Particularly in GCs, one of the most important physical processes leading up to mergers is the torquing of binary’s orbits produced by the SMBH, process known as the von Zeipel-Lidov-Kozai (ZLK) mechanism (Antonini & Perets, 2012; Stephan et al., 2019; Hoang et al., 2019; Petrovich & Antonini, 2017). More generally, we can consider the tidal field from both the SMBH and the cluster which produces a behaviour similar to the ZLK mechanism (Petrovich & Antonini, 2017; Hamilton & Rafikov, 2019; Bub & Petrovich, 2020; Hamilton & Rafikov, 2021).

The strong tidal field due to the SMBH and NSC potentials has demonstrated to play a key role in the evolution of binaries in the GC, as it efficiently torques the orbits of these binaries resulting in near unity eccentricities that lead to mergers. Recent efforts have been made in accounting for different tidal field-driven physical processes in binary evolution such as cluster triaxiality (Bub & Petrovich, 2020) and relativistic phase space diffusion (Hamilton & Rafikov, 2023), however none has focused on combining these processes with others such as close encounters with field stars (flybys). Although the dynamics of wide binaries when under the influence of tidal fields (Petrovich & Antonini, 2017; Hamilton & Rafikov, 2019a; Rasskazov & Rafikov, 2023; Modak & Hamilton, 2023) and flybys (Collins & Sari, 2008; Leigh et al., 2017; Michaely & Perets, 2020; Hamilton & Modak, 2023) have been studied separately, the potential synergy between both effects has yet to be investigated in more depth.

Due to the high velocity dispersion σ𝜎\sigmaitalic_σ in the GC, most binaries in the GC are soft (e.g., σ𝜎\sigmaitalic_σ is larger than their Keplerian velocity) (Stephan et al., 2016; Alexander & Pfuhl, 2014; Rose et al., 2020; Gautam et al., 2024). One single encounter can potentially reset a binary’s orbit, changing its orientation and even pushing it towards extreme eccentricities. The semi-major axis of these binaries will also tend to increase due to these encounters (Heggie, 1975; Hills, 1975a, b). This orbit widening limits the efficiency for tidal field driven mergers, and previous works have simply prescribed it as a limiting evaporation timescale tevapsubscript𝑡evapt_{\text{evap}}italic_t start_POSTSUBSCRIPT evap end_POSTSUBSCRIPT (e.g., Stephan et al. 2016; Petrovich & Antonini 2017; Hoang et al. 2018), using tevapsubscript𝑡evapt_{\text{evap}}italic_t start_POSTSUBSCRIPT evap end_POSTSUBSCRIPT from Binney & Tremaine (2008). However, beyond the orbit softening, flyby interactions can produce a random walk in the binary’s angular momentum j𝑗\vec{j}over→ start_ARG italic_j end_ARG which may lead to mergers as studied by Michaely & Perets (2020). A study looking into the effect from flybys in the cluster-driven dynamics is largely missing in this context (GCs) and the focus of our work.

The presented dynamics is reminiscent of that from wide binaries in general, including comets in the Oort Cloud. In particular, (Heisler & Tremaine, 1986) studied how the galactic tidal field may torque the orbit of binaries and found that flybys can contribute to the diffusion in j𝑗jitalic_j, becoming a considerable contribution to the loss rate of comets in the Oort Cloud. In contrast, as we show in this paper, binaries in GCs have slow diffusion rates compared to the torquing rates that drastically promote the merger fraction.

In this paper we study the binary evolution due to both stellar flybys and tidal fields, proving that by combining both mechanisms can drive these binaries in galactic centers to extreme eccentricities much more efficiently than when these processes are considered separately. The paper is organised as follows. We introduce the basic dynamics for binaries in tidal fields in section §2 and discuss the specific case of the GC including flybys in section §3. We present our results from numerical simulations in section §4 and we discuss the synergy found between tidal fields and flybys in §5, along with the effects on relativistic quenching in section §6. Finally, we present an overall discussion in section §7 and summarize and present the main conclusions in section §8.

2 Dynamical evolution due to tidal fields

As binaries are usually immersed in stellar clusters, an important aspect of their dynamical evolution will be driven by the smooth tidal field of said cluster. These tidal fields drive secular oscillations in a binary’s eccentricity and inclination, similar to the ZLK mechanism. In this section we model this evolution mathematically, following the work done by Hamilton & Rafikov (2019a, b). We will refer to the compact object binary system as the inner binary, while the orbit of this system around the central SMBH will be referred to as the outer binary.

2.1 Equations of motion

When studying orbits in axisymmetric potentials, the evolution of the inner binary can be approximated by averaging the tidal tensors over several periods of the outer binary. This is known as “torus-averaging”; when the outer orbit of the binary is not closed, it will trace a non-repeating path around the cluster which will, after many orbits, densely fill an axisymmetric torus. For any function that follows the outer orbit of the binary, a weighted volume average over this torus can be used instead of a time average.

As shown by Hamilton & Rafikov (2019a), in this torus-averaged limit the tidal tensor of a given potential ΦΦ\Phiroman_Φ only has diagonal terms, and Φxx=Φyydelimited-⟨⟩subscriptΦ𝑥𝑥delimited-⟨⟩subscriptΦ𝑦𝑦\langle\Phi_{xx}\rangle=\langle\Phi_{yy}\rangle⟨ roman_Φ start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT ⟩ = ⟨ roman_Φ start_POSTSUBSCRIPT italic_y italic_y end_POSTSUBSCRIPT ⟩. Here it is useful to introduce Milankovitch’s eccentricity and dimensionless angular momentum vectors {e,j}𝑒𝑗\{\vec{e},\vec{j}\}{ over→ start_ARG italic_e end_ARG , over→ start_ARG italic_j end_ARG }, Milankovitch’s eccentricity and dimensionless angular momentum vectors (e.g., see Tremaine 2023), in order to simplify notation. For a smooth potential ΦΦ\Phiroman_Φ that changes over scales that are much larger than that of the inner binary’s semi-major axis (expanded up to the quadrupole tidal field), this results in the torus-averaged potential

Φ=3ain22Φzz+Φxx[12Γ(5ez2jz2)+14e2(15Γ)],delimited-⟨⟩Φ3superscriptsubscript𝑎in22delimited-⟨⟩subscriptΦ𝑧𝑧subscriptΦ𝑥𝑥delimited-[]12Γ5superscriptsubscript𝑒𝑧2superscriptsubscript𝑗𝑧214superscript𝑒215Γ\langle\Phi\rangle=\frac{3a_{\text{in}}^{2}}{2}\langle\Phi_{zz}+\Phi_{xx}% \rangle\left[\frac{1}{2}\Gamma(5e_{z}^{2}-j_{z}^{2})+\frac{1}{4}e^{2}(1-5% \Gamma)\right],⟨ roman_Φ ⟩ = divide start_ARG 3 italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG ⟨ roman_Φ start_POSTSUBSCRIPT italic_z italic_z end_POSTSUBSCRIPT + roman_Φ start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT ⟩ [ divide start_ARG 1 end_ARG start_ARG 2 end_ARG roman_Γ ( 5 italic_e start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_j start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) + divide start_ARG 1 end_ARG start_ARG 4 end_ARG italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( 1 - 5 roman_Γ ) ] , (1)

where

ΓΦzzΦxx3Φzz+Φxx.Γdelimited-⟨⟩subscriptΦ𝑧𝑧subscriptΦ𝑥𝑥3delimited-⟨⟩subscriptΦ𝑧𝑧subscriptΦ𝑥𝑥\Gamma\equiv\frac{\langle\Phi_{zz}-\Phi_{xx}\rangle}{3\langle\Phi_{zz}+\Phi_{% xx}\rangle}.roman_Γ ≡ divide start_ARG ⟨ roman_Φ start_POSTSUBSCRIPT italic_z italic_z end_POSTSUBSCRIPT - roman_Φ start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT ⟩ end_ARG start_ARG 3 ⟨ roman_Φ start_POSTSUBSCRIPT italic_z italic_z end_POSTSUBSCRIPT + roman_Φ start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT ⟩ end_ARG . (2)

From this potential, a doubly-averaged Hamiltonian can be written in terms of ΓΓ\Gammaroman_Γ as H=HK+H1𝐻subscript𝐻𝐾subscript𝐻1H=H_{K}+H_{1}italic_H = italic_H start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT + italic_H start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT where HKsubscript𝐻𝐾H_{K}italic_H start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT is the Keplerian term and the dimensionless perturbing term is given by (Hamilton & Rafikov, 2019b):

H1=2[132e2(5Γ1)3Γ(jj^out)2+15Γ(ej^out)2].subscript𝐻12delimited-[]132superscript𝑒25Γ13Γsuperscript𝑗subscript^𝑗out215Γsuperscript𝑒subscript^𝑗out2\begin{split}H_{1}=2\Bigl{[}1-\frac{3}{2}e^{2}(5\Gamma-1)-3\Gamma(\vec{j}\cdot% \hat{j}_{\text{out}})^{2}+15\Gamma(\vec{e}\cdot\hat{j}_{\text{out}})^{2}\Bigr{% ]}.\end{split}start_ROW start_CELL italic_H start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 2 [ 1 - divide start_ARG 3 end_ARG start_ARG 2 end_ARG italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( 5 roman_Γ - 1 ) - 3 roman_Γ ( over→ start_ARG italic_j end_ARG ⋅ over^ start_ARG italic_j end_ARG start_POSTSUBSCRIPT out end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 15 roman_Γ ( over→ start_ARG italic_e end_ARG ⋅ over^ start_ARG italic_j end_ARG start_POSTSUBSCRIPT out end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] . end_CELL end_ROW (3)

The associated dimensionless time is τ=t/τsec𝜏𝑡subscript𝜏sec\tau=t/\tau_{\text{sec}}italic_τ = italic_t / italic_τ start_POSTSUBSCRIPT sec end_POSTSUBSCRIPT, where τsecsubscript𝜏sec\tau_{\text{sec}}italic_τ start_POSTSUBSCRIPT sec end_POSTSUBSCRIPT is the secular timescale given by

τsec1=3ain3/22GMbinΦzz+Φxx.superscriptsubscript𝜏sec13superscriptsubscript𝑎in322𝐺subscript𝑀bindelimited-⟨⟩subscriptΦ𝑧𝑧subscriptΦ𝑥𝑥\tau_{\text{sec}}^{-1}=\frac{3a_{\text{in}}^{3/2}}{2\sqrt{GM_{\text{bin}}}}% \langle\Phi_{zz}+\Phi_{xx}\rangle.italic_τ start_POSTSUBSCRIPT sec end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT = divide start_ARG 3 italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 / 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 square-root start_ARG italic_G italic_M start_POSTSUBSCRIPT bin end_POSTSUBSCRIPT end_ARG end_ARG ⟨ roman_Φ start_POSTSUBSCRIPT italic_z italic_z end_POSTSUBSCRIPT + roman_Φ start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT ⟩ . (4)

The equations of motion for the binary that we use throughout this work can be written in a vectorial form as:

dedτ=12(5Γ1)(j×e)5Γ(ej^out)(j×j^out)+Γ(jj^out)(e×j^out),𝑑𝑒𝑑𝜏125Γ1𝑗𝑒5Γ𝑒subscript^𝑗out𝑗subscript^𝑗outΓ𝑗subscript^𝑗out𝑒subscript^𝑗out\begin{split}\frac{d\vec{e}}{d\tau}&=\frac{1}{2}(5\Gamma-1)(\vec{j}\times\vec{% e})-5\Gamma(\vec{e}\cdot\hat{j}_{\text{out}})(\vec{j}\times\hat{j}_{\text{out}% })\\ &\qquad+\Gamma(\vec{j}\cdot\hat{j}_{\text{out}})(\vec{e}\times\hat{j}_{\text{% out}}),\\ \end{split}start_ROW start_CELL divide start_ARG italic_d over→ start_ARG italic_e end_ARG end_ARG start_ARG italic_d italic_τ end_ARG end_CELL start_CELL = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( 5 roman_Γ - 1 ) ( over→ start_ARG italic_j end_ARG × over→ start_ARG italic_e end_ARG ) - 5 roman_Γ ( over→ start_ARG italic_e end_ARG ⋅ over^ start_ARG italic_j end_ARG start_POSTSUBSCRIPT out end_POSTSUBSCRIPT ) ( over→ start_ARG italic_j end_ARG × over^ start_ARG italic_j end_ARG start_POSTSUBSCRIPT out end_POSTSUBSCRIPT ) end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL + roman_Γ ( over→ start_ARG italic_j end_ARG ⋅ over^ start_ARG italic_j end_ARG start_POSTSUBSCRIPT out end_POSTSUBSCRIPT ) ( over→ start_ARG italic_e end_ARG × over^ start_ARG italic_j end_ARG start_POSTSUBSCRIPT out end_POSTSUBSCRIPT ) , end_CELL end_ROW (5)
djdτ=Γ(jj^out)(j×j^out)5Γ(ej^out)(e×j^out),𝑑𝑗𝑑𝜏Γ𝑗subscript^𝑗out𝑗subscript^𝑗out5Γ𝑒subscript^𝑗out𝑒subscript^𝑗out\begin{split}\frac{d\vec{j}}{d\tau}&=\Gamma(\vec{j}\cdot\hat{j}_{\text{out}})(% \vec{j}\times\hat{j}_{\text{out}})-5\Gamma(\vec{e}\cdot\hat{j}_{\text{out}})(% \vec{e}\times\hat{j}_{\text{out}}),\end{split}start_ROW start_CELL divide start_ARG italic_d over→ start_ARG italic_j end_ARG end_ARG start_ARG italic_d italic_τ end_ARG end_CELL start_CELL = roman_Γ ( over→ start_ARG italic_j end_ARG ⋅ over^ start_ARG italic_j end_ARG start_POSTSUBSCRIPT out end_POSTSUBSCRIPT ) ( over→ start_ARG italic_j end_ARG × over^ start_ARG italic_j end_ARG start_POSTSUBSCRIPT out end_POSTSUBSCRIPT ) - 5 roman_Γ ( over→ start_ARG italic_e end_ARG ⋅ over^ start_ARG italic_j end_ARG start_POSTSUBSCRIPT out end_POSTSUBSCRIPT ) ( over→ start_ARG italic_e end_ARG × over^ start_ARG italic_j end_ARG start_POSTSUBSCRIPT out end_POSTSUBSCRIPT ) , end_CELL end_ROW (6)

where we shall assume that j^outsubscript^𝑗out\hat{j}_{\text{out}}over^ start_ARG italic_j end_ARG start_POSTSUBSCRIPT out end_POSTSUBSCRIPT remains fixed throughout the evolution. The latter approximation assumes that either the tidal field is spherically symmetric or the outer orbit lies in the meridional plane of an axisymmetric field (Petrovich & Antonini, 2017).

As discussed by Hamilton & Rafikov (2019b, a), ΓΓ\Gammaroman_Γ determines the phase-space structure of the Hamiltonian, and hence is a useful parameter to classify different dynamical regimes for binary evolution. For example, Γ=1Γ1\Gamma=1roman_Γ = 1 when the potential is Keplerian (only a SMBH at the center), and hence the Hamiltonian in equation (3) reduces to the quadrupole ZLK Hamiltonian. The secular timescale reduces to the well-known ZLK timescale (e.g., Holman et al. 1997):

τZLKsubscript𝜏ZLK\displaystyle\tau_{\rm ZLK}italic_τ start_POSTSUBSCRIPT roman_ZLK end_POSTSUBSCRIPT =\displaystyle== 4aout3(1eout2)3/2Mbin1/23ain3/2MBHG1/24subscriptsuperscript𝑎3outsuperscript1superscriptsubscript𝑒out232subscriptsuperscript𝑀12bin3superscriptsubscript𝑎in32subscript𝑀BHsuperscript𝐺12\displaystyle\frac{4a^{3}_{\text{out}}(1-e_{\text{out}}^{2})^{3/2}M^{1/2}_{% \text{bin}}}{3a_{\text{in}}^{3/2}M_{\text{BH}}G^{1/2}}divide start_ARG 4 italic_a start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT out end_POSTSUBSCRIPT ( 1 - italic_e start_POSTSUBSCRIPT out end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 3 / 2 end_POSTSUPERSCRIPT italic_M start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bin end_POSTSUBSCRIPT end_ARG start_ARG 3 italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 / 2 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT BH end_POSTSUBSCRIPT italic_G start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT end_ARG (7)
similar-to-or-equals\displaystyle\simeq 3×105(Mbin10M)(MBH4×106M)13superscript105subscript𝑀bin10subscript𝑀direct-productsuperscriptsubscript𝑀BH4superscript106subscript𝑀direct-product1\displaystyle 3\times 10^{5}\left(\frac{M_{\rm bin}}{10M_{\odot}}\right)\left(% \frac{M_{\text{BH}}}{4\times 10^{6}M_{\odot}}\right)^{-1}3 × 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT ( divide start_ARG italic_M start_POSTSUBSCRIPT roman_bin end_POSTSUBSCRIPT end_ARG start_ARG 10 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT end_ARG ) ( divide start_ARG italic_M start_POSTSUBSCRIPT BH end_POSTSUBSCRIPT end_ARG start_ARG 4 × 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT
×\displaystyle\times× (ain25au)3/2(aout(1eout2)1/20.3pc)3year.superscriptsubscript𝑎in25au32superscriptsubscript𝑎outsuperscript1subscriptsuperscript𝑒2out120.3pc3year\displaystyle\left(\frac{a_{\text{in}}}{25\rm au}\right)^{-3/2}\left(\frac{a_{% \text{out}}(1-e^{2}_{\text{out}})^{1/2}}{0.3\rm pc}\right)^{3}\text{year}.( divide start_ARG italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT end_ARG start_ARG 25 roman_a roman_u end_ARG ) start_POSTSUPERSCRIPT - 3 / 2 end_POSTSUPERSCRIPT ( divide start_ARG italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT ( 1 - italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT out end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT end_ARG start_ARG 0.3 roman_pc end_ARG ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT year .

We note that the ZLK Hamiltonian assumes the test particle approximation and sets eout=0subscript𝑒out0e_{\text{out}}=0italic_e start_POSTSUBSCRIPT out end_POSTSUBSCRIPT = 0, so that the quadrupole level expansion is a good approximation of the secular dynamics. If, instead, the outer orbit is eccentric, the z𝑧zitalic_z-component of the angular momentum is no longer conserved, and hence the octupole expansion is required (Naoz, 2016). This expansion allows for qualitatively different behavior, where extremely high eccentricities for the inner orbit can be achieved along with orbital flipping (Naoz et al., 2011, 2013; Hoang et al., 2018).

Similarly, we can obtain the Hamiltonian for a binary evolving in a thin Galactic disk by replacing Γ=1/3Γ13\Gamma=1/3roman_Γ = 1 / 3 (see discussion in §7.1), while realistic spherical potentials can be reproduced with 0Γ10Γ10\leqslant\Gamma\leqslant 10 ⩽ roman_Γ ⩽ 1 as shown by Hamilton & Rafikov (2019a).

An important property of the secular Hamiltonian in Equation (3) is its axissymmetry (i.e., invariability to rotations around j^outsubscript^𝑗out\hat{j}_{\text{out}}over^ start_ARG italic_j end_ARG start_POSTSUBSCRIPT out end_POSTSUBSCRIPT ), which implies the z𝑧zitalic_z-component of the angular momentum ain(1ein2)cosiproportional-toabsentsubscript𝑎in1superscriptsubscript𝑒in2𝑖\propto\sqrt{a_{\text{in}}(1-e_{\text{in}}^{2})}\cos i∝ square-root start_ARG italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT ( 1 - italic_e start_POSTSUBSCRIPT in end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_ARG roman_cos italic_i with cosiin=j^inj^outsubscript𝑖insubscript^𝑗insubscript^𝑗out\cos i_{\text{in}}=\hat{j}_{\text{in}}\cdot\hat{j}_{\text{out}}roman_cos italic_i start_POSTSUBSCRIPT in end_POSTSUBSCRIPT = over^ start_ARG italic_j end_ARG start_POSTSUBSCRIPT in end_POSTSUBSCRIPT ⋅ over^ start_ARG italic_j end_ARG start_POSTSUBSCRIPT out end_POSTSUBSCRIPT is conserved. Thus, the Hamiltonian is integrable and we can express the maximum eccentricity due to tidal fields as (Hamilton & Rafikov, 2019a):

emax=110Γcos2i05Γ+1,subscript𝑒max110Γsuperscript2subscript𝑖05Γ1e_{\text{max}}=\sqrt{1-\frac{10\Gamma\cos^{2}{i_{0}}}{5\Gamma+1}},italic_e start_POSTSUBSCRIPT max end_POSTSUBSCRIPT = square-root start_ARG 1 - divide start_ARG 10 roman_Γ roman_cos start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_i start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG 5 roman_Γ + 1 end_ARG end_ARG , (8)

where i0subscript𝑖0i_{0}italic_i start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the initial initial when e02superscriptsubscript𝑒02e_{0}^{2}italic_e start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT is nearly 0.

Finally, in addition to the equations of motion (5) and (6), we add an extra term in Equation (5) to account for the relativistic precession given by

dedt|GR=ω˙GR(1e2)3/2j×e,evaluated-at𝑑𝑒𝑑𝑡GRsubscript˙𝜔GRsuperscript1superscript𝑒232𝑗𝑒\frac{d\vec{e}}{dt}\Big{|}_{\text{GR}}=\frac{\dot{\omega}_{\text{GR}}}{(1-e^{2% })^{3/2}}\vec{j}\times\vec{e},divide start_ARG italic_d over→ start_ARG italic_e end_ARG end_ARG start_ARG italic_d italic_t end_ARG | start_POSTSUBSCRIPT GR end_POSTSUBSCRIPT = divide start_ARG over˙ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT GR end_POSTSUBSCRIPT end_ARG start_ARG ( 1 - italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 3 / 2 end_POSTSUPERSCRIPT end_ARG over→ start_ARG italic_j end_ARG × over→ start_ARG italic_e end_ARG , (9)

where

ω˙GR=3G3/2Mbin3/2ain5/2c2.subscript˙𝜔GR3superscript𝐺32superscriptsubscript𝑀bin32superscriptsubscript𝑎in52superscript𝑐2\dot{\omega}_{\text{GR}}=\frac{3G^{3/2}M_{\text{bin}}^{3/2}}{a_{\text{in}}^{5/% 2}c^{2}}.over˙ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT GR end_POSTSUBSCRIPT = divide start_ARG 3 italic_G start_POSTSUPERSCRIPT 3 / 2 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT bin end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 / 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 5 / 2 end_POSTSUPERSCRIPT italic_c start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG . (10)

2.2 Merger conditions

A major application we consider in the galactic center is the evolution and potential merger of compact object binaries. When the two components of these binaries come very close together, they lose energy by GW radiation. The merger timescale for this gravitational radiation is defined by Peters (1964) as

τGW=385(ain4c5G3m1m2Mbin)(1ein)7/2,subscript𝜏GW385superscriptsubscript𝑎in4superscript𝑐5superscript𝐺3subscript𝑚1subscript𝑚2subscript𝑀binsuperscript1subscript𝑒in72\tau_{\text{GW}}=\frac{3}{85}\left(\frac{a_{\text{in}}^{4}c^{5}}{G^{3}m_{1}m_{% 2}M_{\text{bin}}}\right)(1-e_{\text{in}})^{7/2},italic_τ start_POSTSUBSCRIPT GW end_POSTSUBSCRIPT = divide start_ARG 3 end_ARG start_ARG 85 end_ARG ( divide start_ARG italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT italic_c start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT end_ARG start_ARG italic_G start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_m start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_M start_POSTSUBSCRIPT bin end_POSTSUBSCRIPT end_ARG ) ( 1 - italic_e start_POSTSUBSCRIPT in end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 7 / 2 end_POSTSUPERSCRIPT , (11)

where m1subscript𝑚1m_{1}italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and m2subscript𝑚2m_{2}italic_m start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT If a binary’s orbit shrinks significantly in one secular cycle, then a rapid merger will take place. For this to happen, the timescale for the change in the semi-major axis of the binary due to GW losses (τGWsubscript𝜏GW\tau_{\text{GW}}italic_τ start_POSTSUBSCRIPT GW end_POSTSUBSCRIPT) must be shorter than the timescale for the change in the pericentric distance ain(1ein)subscript𝑎in1subscript𝑒ina_{\text{in}}(1-e_{\text{in}})italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT ( 1 - italic_e start_POSTSUBSCRIPT in end_POSTSUBSCRIPT ) due to tidal fields, which, at high eccentricities, is 1e2τsecsimilar-to-or-equalsabsent1superscript𝑒2subscript𝜏sec\simeq\sqrt{1-e^{2}}\tau_{\rm sec}≃ square-root start_ARG 1 - italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG italic_τ start_POSTSUBSCRIPT roman_sec end_POSTSUBSCRIPT. This results in the condition

1emerger3×105(aout(1eout2)1/20.1pc),less-than-or-similar-to1subscript𝑒merger3superscript105subscript𝑎outsuperscript1superscriptsubscript𝑒out2120.1pc1-e_{\text{merger}}\lesssim 3\times 10^{-5}\left(\frac{a_{\text{out}}(1-e_{% \text{out}}^{2})^{1/2}}{0.1\text{pc}}\right),1 - italic_e start_POSTSUBSCRIPT merger end_POSTSUBSCRIPT ≲ 3 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT ( divide start_ARG italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT ( 1 - italic_e start_POSTSUBSCRIPT out end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT end_ARG start_ARG 0.1 pc end_ARG ) , (12)

for binaries with Mbin=10Msubscript𝑀bin10subscript𝑀direct-productM_{\text{bin}}=10M_{\odot}italic_M start_POSTSUBSCRIPT bin end_POSTSUBSCRIPT = 10 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT initially at ain=10subscript𝑎in10a_{\text{in}}=10italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT = 10au when only ZLK cycles are considered (only SMBH). A similar expression can be obtained for a more general cluster potential (arbitrary ΓΓ\Gammaroman_Γ; see Equation 15 for a secular timescale adding a cluster with a Hernquist potential).

In order for BH (stellar) binaries to reach extremely high eccentricities of 1emax105less-than-or-similar-to1subscript𝑒maxsuperscript1051-e_{\text{max}}\lesssim 10^{-5}1 - italic_e start_POSTSUBSCRIPT max end_POSTSUBSCRIPT ≲ 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT (1emax103less-than-or-similar-to1subscript𝑒maxsuperscript1031-e_{\text{max}}\lesssim 10^{-3}1 - italic_e start_POSTSUBSCRIPT max end_POSTSUBSCRIPT ≲ 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT), we require that ω˙GRτsec0.01less-than-or-similar-tosubscript˙𝜔GRsubscript𝜏sec0.01\dot{\omega}_{\text{GR}}\tau_{\mathrm{sec}}\lesssim 0.01over˙ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT GR end_POSTSUBSCRIPT italic_τ start_POSTSUBSCRIPT roman_sec end_POSTSUBSCRIPT ≲ 0.01 (ω˙GRτsec0.1less-than-or-similar-tosubscript˙𝜔GRsubscript𝜏sec0.1\dot{\omega}_{\text{GR}}\tau_{\mathrm{sec}}\lesssim 0.1over˙ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT GR end_POSTSUBSCRIPT italic_τ start_POSTSUBSCRIPT roman_sec end_POSTSUBSCRIPT ≲ 0.1) as shown by Petrovich & Antonini (2017). In general, we will define extreme eccentricities as 1emax104less-than-or-similar-to1subscript𝑒maxsuperscript1041-e_{\text{max}}\lesssim 10^{-4}1 - italic_e start_POSTSUBSCRIPT max end_POSTSUBSCRIPT ≲ 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT following equation (12), hence binaries that achieve these values will be considered as possible mergers.

3 Dynamics in the Galactic center

In the case of a spherical γ𝛾\gammaitalic_γ-family cluster with a distance scale s𝑠sitalic_s and with a central SMBH in which we have a circular orbit, ΓΓ\Gammaroman_Γ becomes

Γ=3MBH+Mclaout3γ(aout+s)γ4(3aout+sγ)3[MBH+Mclaout3γ(aout+s)γ4(aout+s(4γ))],Γ3subscript𝑀BHsubscript𝑀clsuperscriptsubscript𝑎out3𝛾superscriptsubscript𝑎out𝑠𝛾43subscript𝑎out𝑠𝛾3delimited-[]subscript𝑀BHsubscript𝑀clsuperscriptsubscript𝑎out3𝛾superscriptsubscript𝑎out𝑠𝛾4subscript𝑎out𝑠4𝛾\Gamma=\frac{3M_{\text{BH}}+M_{\text{cl}}a_{\text{out}}^{3-\gamma}(a_{\text{% out}}+s)^{\gamma-4}(3a_{\text{out}}+s\gamma)}{3[M_{\text{BH}}+M_{\text{cl}}a_{% \text{out}}^{3-\gamma}(a_{\text{out}}+s)^{\gamma-4}(a_{\text{out}}+s(4-\gamma)% )]},roman_Γ = divide start_ARG 3 italic_M start_POSTSUBSCRIPT BH end_POSTSUBSCRIPT + italic_M start_POSTSUBSCRIPT cl end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 - italic_γ end_POSTSUPERSCRIPT ( italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT + italic_s ) start_POSTSUPERSCRIPT italic_γ - 4 end_POSTSUPERSCRIPT ( 3 italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT + italic_s italic_γ ) end_ARG start_ARG 3 [ italic_M start_POSTSUBSCRIPT BH end_POSTSUBSCRIPT + italic_M start_POSTSUBSCRIPT cl end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 - italic_γ end_POSTSUPERSCRIPT ( italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT + italic_s ) start_POSTSUPERSCRIPT italic_γ - 4 end_POSTSUPERSCRIPT ( italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT + italic_s ( 4 - italic_γ ) ) ] end_ARG , (13)

where we have considered the tidal averaged tidal tensors Φxxdelimited-⟨⟩subscriptΦ𝑥𝑥\langle\Phi_{xx}\rangle⟨ roman_Φ start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT ⟩ and Φzzdelimited-⟨⟩subscriptΦ𝑧𝑧\langle\Phi_{zz}\rangle⟨ roman_Φ start_POSTSUBSCRIPT italic_z italic_z end_POSTSUBSCRIPT ⟩ as derived by Bub & Petrovich (2020).

In the simplified case that we consider a spherical Hernquist potential with γ=1𝛾1\gamma=1italic_γ = 1 to model the center of our galaxy, this reduces to

Γ=3MBH+Mclaout2(aout+s)3(3aout+s)3[MBH+Mclaout2(aout+s)3(aout+3s)],Γ3subscript𝑀BHsubscript𝑀clsuperscriptsubscript𝑎out2superscriptsubscript𝑎out𝑠33subscript𝑎out𝑠3delimited-[]subscript𝑀BHsubscript𝑀clsuperscriptsubscript𝑎out2superscriptsubscript𝑎out𝑠3subscript𝑎out3𝑠\Gamma=\frac{3M_{\text{BH}}+M_{\text{cl}}a_{\text{out}}^{2}(a_{\text{out}}+s)^% {-3}(3a_{\text{out}}+s)}{3[M_{\text{BH}}+M_{\text{cl}}a_{\text{out}}^{2}(a_{% \text{out}}+s)^{-3}(a_{\text{out}}+3s)]},roman_Γ = divide start_ARG 3 italic_M start_POSTSUBSCRIPT BH end_POSTSUBSCRIPT + italic_M start_POSTSUBSCRIPT cl end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT + italic_s ) start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT ( 3 italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT + italic_s ) end_ARG start_ARG 3 [ italic_M start_POSTSUBSCRIPT BH end_POSTSUBSCRIPT + italic_M start_POSTSUBSCRIPT cl end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT + italic_s ) start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT ( italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT + 3 italic_s ) ] end_ARG , (14)

where s𝑠sitalic_s is the scale radius of the potential.

For small values of aout/ssubscript𝑎out𝑠a_{\text{out}}/sitalic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT / italic_s we have Γ1Γ1\Gamma\approx 1roman_Γ ≈ 1, where the SMBH dominates over the cluster and hence the dynamical evolution of the binary reduces to the ZLK mechanism as mentioned in section §2.1. The same happens when aout/ssubscript𝑎out𝑠a_{\text{out}}/sitalic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT / italic_s becomes very large; the cluster and SMBH in this limit behave as a single body so that as we get further away from the central SMBH, Γ1Γ1\Gamma\rightarrow 1roman_Γ → 1.

The secular timescale in these kinds of environments then becomes (Bub & Petrovich, 2020):

τsec1=3ain3/22GMbin(GMBH2aout3+GMclaoutγ(aout+s)3γ[23aout+sγ2(aout+s)]).superscriptsubscript𝜏sec13superscriptsubscript𝑎in322𝐺subscript𝑀bin𝐺subscript𝑀BH2superscriptsubscript𝑎out3𝐺subscript𝑀clsuperscriptsubscript𝑎out𝛾superscriptsubscript𝑎out𝑠3𝛾delimited-[]23subscript𝑎out𝑠𝛾2subscript𝑎out𝑠\begin{split}\tau_{\text{sec}}^{-1}=&\frac{3a_{\text{in}}^{3/2}}{2\sqrt{GM_{% \text{bin}}}}\Bigg{(}\frac{GM_{\text{BH}}}{2a_{\text{out}}^{3}}\\ &+\frac{GM_{\text{cl}}}{a_{\text{out}}^{\gamma}(a_{\text{out}}+s)^{3-\gamma}}% \left[2-\frac{3a_{\text{out}}+s\gamma}{2(a_{\text{out}}+s)}\right]\Bigg{)}.% \end{split}start_ROW start_CELL italic_τ start_POSTSUBSCRIPT sec end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT = end_CELL start_CELL divide start_ARG 3 italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 / 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 square-root start_ARG italic_G italic_M start_POSTSUBSCRIPT bin end_POSTSUBSCRIPT end_ARG end_ARG ( divide start_ARG italic_G italic_M start_POSTSUBSCRIPT BH end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL + divide start_ARG italic_G italic_M start_POSTSUBSCRIPT cl end_POSTSUBSCRIPT end_ARG start_ARG italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_γ end_POSTSUPERSCRIPT ( italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT + italic_s ) start_POSTSUPERSCRIPT 3 - italic_γ end_POSTSUPERSCRIPT end_ARG [ 2 - divide start_ARG 3 italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT + italic_s italic_γ end_ARG start_ARG 2 ( italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT + italic_s ) end_ARG ] ) . end_CELL end_ROW (15)
Refer to caption
Figure 1: Full evolution of a binary due to both tidal fields and flyby effects. The initial parameters are aout=0.3subscript𝑎out0.3a_{\text{out}}=0.3italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT = 0.3 pc, ain=25subscript𝑎in25a_{\text{in}}=25italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT = 25 AU, ein=0.1subscript𝑒in0.1e_{\text{in}}=0.1italic_e start_POSTSUBSCRIPT in end_POSTSUBSCRIPT = 0.1, ωin=π/4subscript𝜔in𝜋4\omega_{\text{in}}=\pi/4italic_ω start_POSTSUBSCRIPT in end_POSTSUBSCRIPT = italic_π / 4, Ωin=0subscriptΩin0\Omega_{\text{in}}=0roman_Ω start_POSTSUBSCRIPT in end_POSTSUBSCRIPT = 0, iin=80subscript𝑖in80i_{\text{in}}=80italic_i start_POSTSUBSCRIPT in end_POSTSUBSCRIPT = 80°. Each stellar component of the binary has a mass of 5M5subscript𝑀direct-product5M_{\odot}5 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, and we consider the fiducial scenario of MBH=4×106Msubscript𝑀BH4superscript106subscript𝑀direct-productM_{\text{BH}}=4\times 10^{6}M_{\odot}italic_M start_POSTSUBSCRIPT BH end_POSTSUBSCRIPT = 4 × 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT. The red curves represent the evolution in a flyby-free scenario, simply by integrating Equations (5) and (6). In panels (d) and (e), the shaded region contains the median and both upper and lower quartiles.

3.1 Fly-bys

The main focus of this work is to study the additional effect produced by flybys on the evolution of binaries in tidal fields. Due to the high velocity dispersion in the GC, the binaries residing there are mostly soft (vorb<σsubscript𝑣orb𝜎v_{\text{orb}}<\sigmaitalic_v start_POSTSUBSCRIPT orb end_POSTSUBSCRIPT < italic_σ, with σ𝜎\sigmaitalic_σ the velocity dispersion of the cluster), hence we expect flyby encounters to produce a systematic drift in the semi-major axis of the binary. This is a consequence of the Heggie-Hills law, derived using thermodynamic arguments: "hard binaries tend to harden, whereas soft binaries tend to soften" (Hills, 1975a, b; Heggie, 1975).

In order to classify binaries between hard and soft, we can write the velocity dispersion in the galactic center as according to Kocsis & Tremaine (2011)

σ(r)={280kms10.1pc/r10.035(r/0.1pc)2.2,if r<0.22pc250kms10.1pc/r,if r>0.22pc𝜎𝑟cases280superscriptkms10.1pc𝑟10.035superscript𝑟0.1pc2.2otherwiseif 𝑟0.22pcotherwise250superscriptkms10.1pc𝑟if 𝑟0.22pcotherwise\sigma(r)=\begin{cases}280\text{kms}^{-1}\sqrt{0.1\text{pc}/r}\sqrt{1-0.035(r/% 0.1\text{pc})^{2.2}},\\ \phantom{280\text{kms}^{-1}\sqrt{0.1\text{pc}/r}\sqrt{1-0.03}}\text{if }r<0.22% \text{pc}\\ 250\text{kms}^{-1}\sqrt{0.1\text{pc}/r},\phantom{\sqrt{01\text{pc}/r}}\text{if% }r>0.22\text{pc}\end{cases}italic_σ ( italic_r ) = { start_ROW start_CELL 280 kms start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT square-root start_ARG 0.1 pc / italic_r end_ARG square-root start_ARG 1 - 0.035 ( italic_r / 0.1 pc ) start_POSTSUPERSCRIPT 2.2 end_POSTSUPERSCRIPT end_ARG , end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL if italic_r < 0.22 pc end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL 250 kms start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT square-root start_ARG 0.1 pc / italic_r end_ARG , if italic_r > 0.22 pc end_CELL start_CELL end_CELL end_ROW (16)

The hard-soft boundary then is given by vkep=σsubscript𝑣kep𝜎v_{\rm{kep}}=\sigmaitalic_v start_POSTSUBSCRIPT roman_kep end_POSTSUBSCRIPT = italic_σ. For example, if we consider a twin binary with two 5M5subscript𝑀direct-product5M_{\odot}5 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT components at aout=0.3subscript𝑎out0.3a_{\text{out}}=0.3italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT = 0.3pc , this boundary would be given by ain=0.2subscript𝑎in0.2a_{\text{in}}=0.2italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT = 0.2au. If we increase the mass of the binary component this limit will occur at a larger value; for 10M10subscript𝑀direct-product10M_{\odot}10 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT components it occurs at ain=0.4subscript𝑎in0.4a_{\text{in}}=0.4italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT = 0.4au, and for 50M50subscript𝑀direct-product50M_{\odot}50 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT at ain=2.1subscript𝑎in2.1a_{\text{in}}=2.1italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT = 2.1au. These continue to be relatively small values, further reinforcing that most binaries in the galactic center can be classified as soft.

The time between flyby encounters

tenc=(nσpvp)1,subscript𝑡encsuperscriptsubscript𝑛subscript𝜎𝑝subscript𝑣𝑝1t_{\rm enc}=(n_{\star}\sigma_{p}v_{p})^{-1},italic_t start_POSTSUBSCRIPT roman_enc end_POSTSUBSCRIPT = ( italic_n start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT , (17)

where σp=πbmax2subscript𝜎𝑝𝜋superscriptsubscript𝑏max2\sigma_{p}=\pi b_{\text{max}}^{2}italic_σ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = italic_π italic_b start_POSTSUBSCRIPT max end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT is the geometric cross section considered for the flybys and nsubscript𝑛n_{\star}italic_n start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT is the stellar number density of the environment. For the galactic center it is considered as n106pc3subscript𝑛superscript106superscriptpc3n_{\star}\approx 10^{6}\text{pc}^{-3}italic_n start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT ≈ 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT pc start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT (Hamers & Tremaine, 2017) and n0.1pc3subscript𝑛0.1superscriptpc3n_{\star}\approx 0.1\text{pc}^{-3}italic_n start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT ≈ 0.1 pc start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT for the solar neighbourhood. It can be more accuratelly modelled by (Kocsis & Tremaine, 2011)

n(r)=2.8×106Mpc3m(r0.22pc)γ,subscript𝑛𝑟2.8superscript106subscriptMdirect-productsuperscriptpc3delimited-⟨⟩subscript𝑚superscript𝑟0.22pc𝛾n_{\star}(r)=\frac{2.8\times 10^{6}\text{M}_{\odot}\text{pc}^{-3}}{\langle m_{% \star}\rangle}\left(\frac{r}{0.22\text{pc}}\right)^{-\gamma},italic_n start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT ( italic_r ) = divide start_ARG 2.8 × 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT pc start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT end_ARG start_ARG ⟨ italic_m start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT ⟩ end_ARG ( divide start_ARG italic_r end_ARG start_ARG 0.22 pc end_ARG ) start_POSTSUPERSCRIPT - italic_γ end_POSTSUPERSCRIPT , (18)

with γ=1.2𝛾1.2\gamma=1.2italic_γ = 1.2 inside 0.22pc and γ=1.75𝛾1.75\gamma=1.75italic_γ = 1.75 outside 0.22pc. We take m1Mdelimited-⟨⟩subscript𝑚1subscript𝑀direct-product\langle m_{\star}\rangle\approx 1M_{\odot}⟨ italic_m start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT ⟩ ≈ 1 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT as the average stellar mass in the GC; Alexander & Pfuhl (2013) presents equations that result in mdelimited-⟨⟩subscript𝑚\langle m_{\star}\rangle⟨ italic_m start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT ⟩ ranging from 1M1subscript𝑀direct-product1M_{\odot}1 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT to 2M2subscript𝑀direct-product2M_{\odot}2 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, therefore to simplify calculations we take all flybys to have a solar mass.

4 Numerical simulations

Because the interaction time due to a flyby (order ain/σsubscript𝑎in𝜎a_{\text{in}}/\sigmaitalic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT / italic_σ, 0.2similar-toabsent0.2\sim 0.2∼ 0.2 years) is much shorter than the secular timescale, we evolve an individual binary using the secular equation of motions in (5) and (6) subject to discrete instances where a flyby takes place, modifying the orbital elements of the binary, and continue the secular evolution until the next flyby takes place. In other words, we switch back and forth between tidal fields and flybys.

The flybys are modelled using direct N-body integrations with REBOUND (Rein & Liu, 2012), with an encounter rate as specified in equation (17) defining bmax=5ainsubscript𝑏max5subscript𝑎inb_{\text{max}}=5a_{\text{in}}italic_b start_POSTSUBSCRIPT max end_POSTSUBSCRIPT = 5 italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT to avoid considering flybys that are too distant. This definition allows us to stay close to the regime in which the impulse approximation is valid, which will prove to be relevant in section §5. We repeated our fiducial numerical experiments in which we doubled bmaxsubscript𝑏maxb_{\text{max}}italic_b start_POSTSUBSCRIPT max end_POSTSUBSCRIPT to 10ain10subscript𝑎in10a_{\text{in}}10 italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT, and found that the outcome was statically similar to those with bmax=5ainsubscript𝑏max5subscript𝑎inb_{\text{max}}=5a_{\text{in}}italic_b start_POSTSUBSCRIPT max end_POSTSUBSCRIPT = 5 italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT, plus the fact that a higher value for bmaxsubscript𝑏maxb_{\text{max}}italic_b start_POSTSUBSCRIPT max end_POSTSUBSCRIPT implies a higher frequency for flyby encounters, which is much more computationally expensive to simulate.

We establish the velocity of our flybys to be constant at vp=200subscript𝑣𝑝200v_{p}=200italic_v start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = 200 kms-1, considering that within 1 pc of the center of our galaxy σ𝜎\sigmaitalic_σ varies roughly between 100 and 300 kms-1 as according to equation (16). These simplifications allow for controlled simulations, which makes comparisons with analytical estimates such as the ones made in section §5 easier.

Figure 1 shows the complete dynamical evolution of an individual binary with aout=0.3subscript𝑎out0.3a_{\text{out}}=0.3italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT = 0.3pc. The red curves show the evolution of an identical binary in a flyby-free scenario, where we only get tidal field driven evolution. It is quite clear by comparing these curves with our results that the inclusion of flybys in our model results in a more dramatic evolution which involves extreme eccentricities, orbital flips and jumps between the einωinsubscript𝑒insubscript𝜔ine_{\text{in}}-\omega_{\text{in}}italic_e start_POSTSUBSCRIPT in end_POSTSUBSCRIPT - italic_ω start_POSTSUBSCRIPT in end_POSTSUBSCRIPT phase-space curves.

The individual changes in ainsubscript𝑎ina_{\text{in}}italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT presented in panel (a) follow a random walk, but show an overall tendency of making the binary wider. This is expected due to the Heggie-Hills law, as mentioned earlier. We can also easily see in panel (b) that by only taking into account tidal field driven evolution, such high eccentricities would never have been achieved by this particular binary. However, this becomes possible when including flybys in the model, which allows the binary to reach values of 1ein1051subscript𝑒insuperscript1051-e_{\text{in}}\approx 10^{-5}1 - italic_e start_POSTSUBSCRIPT in end_POSTSUBSCRIPT ≈ 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT. As discussed in section §2.2, with these extreme eccentricities a rapid merger can be produced due to GW tidal captures.

It is also interesting to note that in panel (c) we can see orbital flips, in which the orbit switches between prograde and retrograde. Considering that the Hamiltonian shown in Equation (3) is only of quadrupole order and assumes eout=0subscript𝑒out0e_{\text{out}}=0italic_e start_POSTSUBSCRIPT out end_POSTSUBSCRIPT = 0, these flips should not be possible as jz=1ein2cosiinsubscript𝑗𝑧1superscriptsubscript𝑒in2subscript𝑖inj_{z}=\sqrt{1-e_{\text{in}}^{2}}\cos i_{\text{in}}italic_j start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = square-root start_ARG 1 - italic_e start_POSTSUBSCRIPT in end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG roman_cos italic_i start_POSTSUBSCRIPT in end_POSTSUBSCRIPT is fixed (e.g., Naoz 2016). However, as flybys modify both einsubscript𝑒ine_{\text{in}}italic_e start_POSTSUBSCRIPT in end_POSTSUBSCRIPT and iinsubscript𝑖ini_{\text{in}}italic_i start_POSTSUBSCRIPT in end_POSTSUBSCRIPT driving a slow random walk in jzsubscript𝑗𝑧j_{z}italic_j start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT (see also the examples in Figure 5), the orbit is allowed to flip, a phenomenon observed during the high eccentricity episodes.

Panels (d) and (e) show the variation in ωinsubscript𝜔in\omega_{\text{in}}italic_ω start_POSTSUBSCRIPT in end_POSTSUBSCRIPT and iinsubscript𝑖ini_{\text{in}}italic_i start_POSTSUBSCRIPT in end_POSTSUBSCRIPT product of each flyby encounter. We can see that both these variations in general are very small, however, along with the variations in eccentricity due to each encounter, they are enough to make the binary’s orbit jump from one phase-space curve to another as is shown in panel (f). Since flybys not only affect the magnitude of the eccentricity but also the orientation of the orbit, we get changes both in ωinsubscript𝜔in\omega_{\text{in}}italic_ω start_POSTSUBSCRIPT in end_POSTSUBSCRIPT and in einsubscript𝑒ine_{\text{in}}italic_e start_POSTSUBSCRIPT in end_POSTSUBSCRIPT which produce these ‘jumps’ in the phase-space portrait. In particular, in a flyby-free scenario the orbit would librate around the fixed point ωin=π/2subscript𝜔in𝜋2\omega_{\text{in}}=\pi/2italic_ω start_POSTSUBSCRIPT in end_POSTSUBSCRIPT = italic_π / 2, however by including flybys in our model the orbit generally circulates one as shown in panel (f).

4.1 Population statistics

Refer to caption
Figure 2: CDFs of the maximum eccentricities reached in the galactic center after 10, 25, 50 and 100 initial secular timescales τ0subscript𝜏0\tau_{0}italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, as labeled. The dashed histogram represents the CDF for a flyby free population, evolved for 100τ0100subscript𝜏0100\tau_{0}100 italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT.

In order to quantify the overall effects of tidal fields and flybys on binary evolution, we generated a series of simplified population synthesis models. It is important to note that all binaries within the population are independent from one another, hence we are not considering possible effects due to binary-binary interactions. We also consider a dynamical stability criterion, so as to not consider binaries that become extremely wide and hence unstable when (Grishin et al., 2016)

ain>0.25aout(MbinMBH)1/3.subscript𝑎in0.25subscript𝑎outsuperscriptsubscript𝑀binsubscript𝑀BH13a_{\text{in}}>0.25a_{\text{out}}\left(\frac{M_{\text{bin}}}{M_{\text{BH}}}% \right)^{1/3}.italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT > 0.25 italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT ( divide start_ARG italic_M start_POSTSUBSCRIPT bin end_POSTSUBSCRIPT end_ARG start_ARG italic_M start_POSTSUBSCRIPT BH end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT 1 / 3 end_POSTSUPERSCRIPT . (19)

In Figure 2 we show the CDF of the maximum eccentricities achieved by a population of 700similar-toabsent700\sim 700∼ 700 binaries in the galactic center at different stages of their evolution. The initial eccentricities of these binaries follow a thermal distribution such that f(e)de=2ede𝑓𝑒𝑑𝑒2𝑒𝑑𝑒f(e)de=2edeitalic_f ( italic_e ) italic_d italic_e = 2 italic_e italic_d italic_e, and their initial inclinations are uniform such that cosi0subscript𝑖0\cos{i_{0}}roman_cos italic_i start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is also uniform between -1 and 1. The eccentricity of the outer orbits is set to 0. All binaries are twins with a total mass of 10M10subscript𝑀direct-product10M_{\odot}10 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT and are evolved for 100τ0100subscript𝜏0100\tau_{0}100 italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, τ0subscript𝜏0\tau_{0}italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT being the initial secular timescale of each individual binary according to equation (15). We consider the fiducial scenario of our GC, where MBH=4×106Msubscript𝑀BH4superscript106subscript𝑀direct-productM_{\text{BH}}=4\times 10^{6}M_{\odot}italic_M start_POSTSUBSCRIPT BH end_POSTSUBSCRIPT = 4 × 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, Mcl8MBHsubscript𝑀cl8subscript𝑀BHM_{\text{cl}}\approx 8M_{\text{BH}}italic_M start_POSTSUBSCRIPT cl end_POSTSUBSCRIPT ≈ 8 italic_M start_POSTSUBSCRIPT BH end_POSTSUBSCRIPT, and s=4𝑠4s=4italic_s = 4 pc. The initial values for both ainsubscript𝑎ina_{\text{in}}italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT and aoutsubscript𝑎outa_{\text{out}}italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT are drawn from a uniform distribution, ranging between 1550155015-5015 - 50au for ainsubscript𝑎ina_{\text{in}}italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT and 0.110.110.1-10.1 - 1pc for aoutsubscript𝑎outa_{\text{out}}italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT. This means that the total population of binaries is initially soft, and that the velocity dispersion does not exceed 275275275275kms-1 making our assumption of vp=200subscript𝑣𝑝200v_{p}=200italic_v start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = 200kms-1 appropriate.

After only a few secular timescales, the initial thermal eccentricity distribution is modified into a wider distribution, achieving higher eccentricities as expected from the tidal fields. We can also observe the cumulative nature of this type of dynamical evolution; the longer we let the binaries evolve, the higher the probability of them reaching extreme eccentricities. This is expected from a chaotic nature of the binary evolution depicted in Figure 1.

Refer to caption
Figure 3: CDFs of the maximum eccentricities and minimum pericenter distances qin=ain(1ein)subscript𝑞insubscript𝑎in1subscript𝑒inq_{\text{in}}=a_{\text{in}}(1-e_{\text{in}})italic_q start_POSTSUBSCRIPT in end_POSTSUBSCRIPT = italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT ( 1 - italic_e start_POSTSUBSCRIPT in end_POSTSUBSCRIPT ) reached in the galactic center after 100 initial secular timescales τ0subscript𝜏0\tau_{0}italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT for different dynamical evolution mechanisms: tidal fields, flybys and a combination of both.
Refer to caption
Figure 4: CDFs of the maximum eccentricities achieved by diffusion in different orbital parameters, evolved for 50τ050subscript𝜏050\tau_{0}50 italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. The black dashed curve represents evolution considering diffusion in all the orbital parameters due to flybys; it is the same as the yellow curve shown in figure 2.
Refer to caption
Figure 5: Evolution of orbital parameters for three binaries with different values of \mathcal{R}caligraphic_R, all with the same initial conditions of ain=25subscript𝑎in25a_{\text{in}}=25italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT = 25au, aout=0.3subscript𝑎out0.3a_{\text{out}}=0.3italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT = 0.3pc, einsubscript𝑒ine_{\text{in}}italic_e start_POSTSUBSCRIPT in end_POSTSUBSCRIPT=0.1 and eout=0subscript𝑒out0e_{\text{out}}=0italic_e start_POSTSUBSCRIPT out end_POSTSUBSCRIPT = 0. All were evolved for 50τ050subscript𝜏050\tau_{0}50 italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, τ0subscript𝜏0\tau_{0}italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT being the initial secular timescale for each binary. Note the difference in the scale of the vertical axis in the top row; as \mathcal{R}caligraphic_R increases, ainsubscript𝑎ina_{\text{in}}italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT grows more drastically.

In Figure 3 we can see the CDFs of the maximum eccentricities achieved after 100τ0100subscript𝜏0100\tau_{0}100 italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT by three different populations of binaries. One population was evolved considering only tidal field effects, the second considering only flybys and the third taking into account both effects (the same population as shown in figure 2). The figure shows that tidal fields are the least efficient dynamical mechanism in producing extreme eccentricities; only about 1%similar-toabsentpercent1\sim 1\%∼ 1 % of binaries evolved with tidal fields alone reach 1emax<1041subscript𝑒maxsuperscript1041-e_{\text{max}}<10^{-4}1 - italic_e start_POSTSUBSCRIPT max end_POSTSUBSCRIPT < 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT. Flybys on the other hand prove to be slightly more effective, driving about 3%percent33\%3 % of binaries in the population to said values. However, when combining both mechanisms a clear synergy is found, increasing this probability up to 34%similar-toabsentpercent34\sim 34\%∼ 34 %. For reference, the bottom panel shows the minimum pericenter distances for the same population of binaries. Although several binaries reach very high eccentricities, as was mentioned in section §3.1 these binaries will tend to get softer, which means that ainsubscript𝑎ina_{\text{in}}italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT and hence the pericentric distance q=ain(1ein)𝑞subscript𝑎in1subscript𝑒inq=a_{\text{in}}(1-e_{\text{in}})italic_q = italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT ( 1 - italic_e start_POSTSUBSCRIPT in end_POSTSUBSCRIPT ) will grow as well. This is where the stability criterion mentioned previously in equation (19) becomes important. As can be seen in the CDF, the extreme values of the eccentricities achieved seem to dominate over the increase in ainsubscript𝑎ina_{\text{in}}italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT, producing a considerable amount of binaries with pericentric distances smaller than 103similar-toabsentsuperscript103\sim 10^{-3}∼ 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPTau, equivalent to a merger timescale τGW<3×105subscript𝜏GW3superscript105\tau_{\rm GW}<3\times 10^{5}italic_τ start_POSTSUBSCRIPT roman_GW end_POSTSUBSCRIPT < 3 × 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT yrs (Eq. [11]). The latter may be shorter than the ZLK timescale in Equation (7) and, therefore, expected to lead to a merger within one secular cycle.

5 Tides vs. flybys

Refer to caption
Figure 6: Merger fraction as a function of the diffusion parameter \mathcal{R}caligraphic_R. The dashed lines indicate the merger fractions for the limits =00\mathcal{R}=0caligraphic_R = 0 (tidal fields alone) and =\mathcal{R}=\inftycaligraphic_R = ∞ (flybys alone). The error bars are evaluated assuming a Poisson distribution as N𝑁\sqrt{N}square-root start_ARG italic_N end_ARG, N𝑁Nitalic_N being the amount of mergers for each bin.
Refer to caption
Figure 7: Evolution of a binary that begins in a GR quenched regime in which mergers are forbidden (colored region). Flyby interactions allow this binary to become softer and move out of this regime, hence becoming a merger candidate. The top panel shows the evolution of ϵGRsubscriptitalic-ϵGR\epsilon_{\text{GR}}italic_ϵ start_POSTSUBSCRIPT GR end_POSTSUBSCRIPT as according to equation (9), and the bottom panel shows the evolution of ainsubscript𝑎ina_{\text{in}}italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT and qinsubscript𝑞inq_{\text{in}}italic_q start_POSTSUBSCRIPT in end_POSTSUBSCRIPT.
Refer to caption
Figure 8: Contours of the diffusion parameter \mathcal{R}caligraphic_R in Equation (22) as a function of ainsubscript𝑎ina_{\text{in}}italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT and aoutsubscript𝑎outa_{\text{out}}italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT, considering the fiducial scenario of MBH=4×106Msubscript𝑀BH4superscript106subscript𝑀direct-productM_{\text{BH}}=4\times 10^{6}M_{\odot}italic_M start_POSTSUBSCRIPT BH end_POSTSUBSCRIPT = 4 × 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, Mp=1Msubscript𝑀𝑝1subscript𝑀direct-productM_{p}=1M_{\odot}italic_M start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = 1 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT and Mbin=10Msubscript𝑀bin10subscript𝑀direct-productM_{\text{bin}}=10M_{\odot}italic_M start_POSTSUBSCRIPT bin end_POSTSUBSCRIPT = 10 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT. The boundary for unstable binaries is plotted as according to Equation (19), while the limit for hard binaries is defined by vkep>σsubscript𝑣kep𝜎v_{\text{kep}}>\sigmaitalic_v start_POSTSUBSCRIPT kep end_POSTSUBSCRIPT > italic_σ. We also plot the region in which relativistic quenching (ϵGR>0.1subscriptitalic-ϵGR0.1\epsilon_{\text{GR}}>0.1italic_ϵ start_POSTSUBSCRIPT GR end_POSTSUBSCRIPT > 0.1) does not allow for mergers.

In this section we analyze in more depth the diffusive effects of flybys on the binary evolution.

As discussed in the previous section, the flybys impart random changes in all orbital elements (eiω𝑒𝑖𝜔e-i-\omegaitalic_e - italic_i - italic_ω), but it remains unclear whether the diffusion in emaxsubscript𝑒maxe_{\rm max}italic_e start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT is dominated by the behavior of a single element or it is a collective process. As a controlled experiment, we show in Figure 4 the CDFs of the maximum eccentricities achieved by populations in which flyby encounters were replaced by random changes in individual orbital elements using the statistical distribution from N-body experiments. We clearly observe that diffusion in the inclination is the main effect driving binaries to high eccentricities. This translates into diffusion of jzsubscript𝑗𝑧j_{z}italic_j start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT as the dominant driver of maximum eccentricity grows. In turn, the diffusion of ωinsubscript𝜔in\omega_{\text{in}}italic_ω start_POSTSUBSCRIPT in end_POSTSUBSCRIPT and einsubscript𝑒ine_{\text{in}}italic_e start_POSTSUBSCRIPT in end_POSTSUBSCRIPT adds little compared to the case where no diffusion occurs (i.e., tidal fields only).

Guided by these numerical experiments, we define a parameter 𝒟𝒟\mathcal{D}caligraphic_D that quantifies the average diffusion rate of the specific angular momentum vector due to a single flyby encounter as

𝒟=Δj2(f,b,i,ω,Ω)tenc.𝒟subscriptdelimited-⟨⟩superscriptnormΔ𝑗2𝑓𝑏𝑖𝜔Ωsubscript𝑡enc\mathcal{D}=\frac{\langle\|\Delta\vec{j}\|^{2}\rangle_{(f,b,i,\omega,\Omega)}}% {t_{\text{enc}}}.caligraphic_D = divide start_ARG ⟨ ∥ roman_Δ over→ start_ARG italic_j end_ARG ∥ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT ( italic_f , italic_b , italic_i , italic_ω , roman_Ω ) end_POSTSUBSCRIPT end_ARG start_ARG italic_t start_POSTSUBSCRIPT enc end_POSTSUBSCRIPT end_ARG . (20)

This coefficient not only includes changes in the specific angular momentum j=1e2𝑗1superscript𝑒2j=\sqrt{1-e^{2}}italic_j = square-root start_ARG 1 - italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG, but also changes in its orientation which is relevant to model the changes in jzsubscript𝑗𝑧j_{z}italic_j start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT. As explained in Appendix A.1, the delimited-⟨⟩\langle\cdot\rangle⟨ ⋅ ⟩ average is carried over an ensemble of binaries with random orientations and phases relative to the stellar perturber.

From this diffusion coefficient we define a dimensionless parameter \mathcal{R}caligraphic_R that quantifies the expected diffusion in j𝑗\vec{j}over→ start_ARG italic_j end_ARG due to flybys after a secular timescale τ0subscript𝜏0\tau_{0}italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT has occurred as

=𝒟τ0.𝒟subscript𝜏0\mathcal{R}=\sqrt{\mathcal{D}\tau_{0}}.caligraphic_R = square-root start_ARG caligraphic_D italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG . (21)

Thus, the limits 00\mathcal{R}\rightarrow 0caligraphic_R → 0 and \mathcal{R}\rightarrow\inftycaligraphic_R → ∞ refer to scenarios completely dominated by tidal fields and by flybys, respectively. In Appendix A.1 we show the full derivation of 𝒟𝒟\mathcal{D}caligraphic_D, in which we have used the impulse approximation to find an analytical expression for ΔjΔ𝑗\Delta\vec{j}roman_Δ over→ start_ARG italic_j end_ARG. From this we obtain

0.02(Mbin10M)14(MpM)(MBH4×106M)12×(vp200km/s)12(ain25au)34(aout0.3pc)32×(n106pc3)12(f(β)1.8)12,similar-to-or-equals0.02superscriptsubscript𝑀bin10subscript𝑀direct-product14subscript𝑀𝑝subscript𝑀direct-productsuperscriptsubscript𝑀BH4superscript106subscript𝑀direct-product12superscriptsubscript𝑣𝑝200km/s12superscriptsubscript𝑎in25au34superscriptsubscript𝑎out0.3pc32superscriptsubscript𝑛superscript106superscriptpc312superscript𝑓𝛽1.812\begin{split}\mathcal{R}\simeq&0.02\left(\frac{M_{\text{bin}}}{10M_{\odot}}% \right)^{-\frac{1}{4}}\left(\frac{M_{p}}{M_{\odot}}\right)\left(\frac{M_{\text% {BH}}}{4\times 10^{6}M_{\odot}}\right)^{-\frac{1}{2}}\\ &\times\left(\frac{v_{p}}{200\text{km/s}}\right)^{-\frac{1}{2}}\left(\frac{a_{% \text{in}}}{25\text{au}}\right)^{\frac{3}{4}}\left(\frac{a_{\text{out}}}{0.3% \text{pc}}\right)^{\frac{3}{2}}\\ &\times\left(\frac{n_{\star}}{10^{6}\text{pc}^{-3}}\right)^{\frac{1}{2}}\left(% \frac{f(\beta)}{1.8}\right)^{\frac{1}{2}},\end{split}start_ROW start_CELL caligraphic_R ≃ end_CELL start_CELL 0.02 ( divide start_ARG italic_M start_POSTSUBSCRIPT bin end_POSTSUBSCRIPT end_ARG start_ARG 10 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT - divide start_ARG 1 end_ARG start_ARG 4 end_ARG end_POSTSUPERSCRIPT ( divide start_ARG italic_M start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG start_ARG italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT end_ARG ) ( divide start_ARG italic_M start_POSTSUBSCRIPT BH end_POSTSUBSCRIPT end_ARG start_ARG 4 × 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT - divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL × ( divide start_ARG italic_v start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG start_ARG 200 km/s end_ARG ) start_POSTSUPERSCRIPT - divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT ( divide start_ARG italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT end_ARG start_ARG 25 au end_ARG ) start_POSTSUPERSCRIPT divide start_ARG 3 end_ARG start_ARG 4 end_ARG end_POSTSUPERSCRIPT ( divide start_ARG italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT end_ARG start_ARG 0.3 pc end_ARG ) start_POSTSUPERSCRIPT divide start_ARG 3 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL × ( divide start_ARG italic_n start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT pc start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT end_ARG ) start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT ( divide start_ARG italic_f ( italic_β ) end_ARG start_ARG 1.8 end_ARG ) start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT , end_CELL end_ROW (22)

where β=s/aout𝛽𝑠subscript𝑎out\beta=s/a_{\text{out}}italic_β = italic_s / italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT and f(β)𝑓𝛽f(\beta)italic_f ( italic_β ) is defined in equation (A8). The accuracy of this expression was verified numerically, by simulating the diffusion in a binary’s angular momentum vector due to only flyby interactions using REBOUND. Figure A2 shows that the numerical equivalent for \mathcal{R}caligraphic_R also converges to 0.020.020.020.02 when considering the fiducial scenario of Mbin=10Msubscript𝑀bin10subscript𝑀direct-productM_{\text{bin}}=10M_{\odot}italic_M start_POSTSUBSCRIPT bin end_POSTSUBSCRIPT = 10 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, Mp=Msubscript𝑀𝑝subscript𝑀direct-productM_{p}=M_{\odot}italic_M start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, MBH=4×106Msubscript𝑀BH4superscript106subscript𝑀direct-productM_{\text{BH}}=4\times 10^{6}M_{\odot}italic_M start_POSTSUBSCRIPT BH end_POSTSUBSCRIPT = 4 × 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, ain=25subscript𝑎in25a_{\text{in}}=25italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT = 25au and aout=0.3subscript𝑎out0.3a_{\text{out}}=0.3italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT = 0.3pc.

In general, we expect different types of evolution when varying \mathcal{R}caligraphic_R, as it should become much more diffusive and incoherent as this parameter grows. Figure 5 shows this behavior, where we have evolved three binaries with different values of \mathcal{R}caligraphic_R by varying MBHsubscript𝑀BHM_{\text{BH}}italic_M start_POSTSUBSCRIPT BH end_POSTSUBSCRIPT in each case. All three binaries have initial parameters of ain=25subscript𝑎in25a_{\text{in}}=25italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT = 25au, aout=0.3subscript𝑎out0.3a_{\text{out}}=0.3italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT = 0.3pc, ein=0.1subscript𝑒in0.1e_{\text{in}}=0.1italic_e start_POSTSUBSCRIPT in end_POSTSUBSCRIPT = 0.1 and eout=0subscript𝑒out0e_{\text{out}}=0italic_e start_POSTSUBSCRIPT out end_POSTSUBSCRIPT = 0, and were evolved for 50τ050subscript𝜏050\tau_{0}50 italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT where τ0subscript𝜏0\tau_{0}italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the initial secular timescale for each case according to equation (15). It is important to note that \mathcal{R}caligraphic_R increases with τsecsubscript𝜏sec\tau_{\text{sec}}italic_τ start_POSTSUBSCRIPT sec end_POSTSUBSCRIPT, therefore in a more diffusive scenario we will have more flybys per secular cycle as the cycles grow longer. We can see in the rightmost column how coherent cycles due to tidal fields become much more difficult to identify when =0.060.06\mathcal{R}=0.06caligraphic_R = 0.06, as opposed to the case where flybys start to become less important as shown in the first column where =0.0060.006\mathcal{R}=0.006caligraphic_R = 0.006. Panel (d) more specifically shows a much less chaotic evolution in which we can easily distinguish coherent cycles with some perturbations such as extreme eccentricities and variations in the secular timescale, as opposed to panel (f) where the cycles are greatly perturbed by the flybys and hence such large eccentricities are not achieved.

The bottom row shows the zlimit-from𝑧z-italic_z -component of the angular momentum jzsubscript𝑗𝑧j_{z}italic_j start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT which should be a conserved quantity in a flyby-free scenario, however we can see how its value drifts due to the diffusion caused by flybys, especially as \mathcal{R}caligraphic_R increases. Due to the increasingly chaotic nature of the evolution of binaries in more diffusive regimes, it becomes much harder to reach the extreme eccentricities required to produce a merger, as can be seen in the case for =0.060.06\mathcal{R}=0.06caligraphic_R = 0.06. The large amount of flybys per secular timescale constantly resets the orbital parameters of the binary, preventing it from completing regular eccentricity cycles and hence decreasing the chances of achieving eccentricities of 1e104less-than-or-similar-to1𝑒superscript1041-e\lesssim 10^{-4}1 - italic_e ≲ 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT.

5.1 Experiments changing \mathcal{R}caligraphic_R

In Figure 6 we show the merger fractions of different binary populations as a function of \mathcal{R}caligraphic_R, which was varied by changing the value of MBHsubscript𝑀BHM_{\text{BH}}italic_M start_POSTSUBSCRIPT BH end_POSTSUBSCRIPT. These populations were all integrated for 15similar-toabsent15\sim 15∼ 15 Myr, which is roughly 50τsec50subscript𝜏sec50\tau_{\text{sec}}50 italic_τ start_POSTSUBSCRIPT sec end_POSTSUBSCRIPT for a binary starting with ain=25subscript𝑎in25a_{\text{in}}=25italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT = 25au at 0.30.30.30.3pc considering the fiducial case of =0.020.02\mathcal{R}=0.02caligraphic_R = 0.02. We consider any binary that reaches 1emax1041subscript𝑒maxsuperscript1041-e_{\text{max}}\leq 10^{-4}1 - italic_e start_POSTSUBSCRIPT max end_POSTSUBSCRIPT ≤ 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT to be a merger, as was discussed in section §2.2. As mentioned previously, a less diffusive regime comes with shorter secular timescales; therefore, by integrating a population with =0.0060.006\mathcal{R}=0.006caligraphic_R = 0.006 for 15151515Myr we are actually letting them evolve for 500τsecsimilar-toabsent500subscript𝜏sec\sim 500\tau_{\text{sec}}∼ 500 italic_τ start_POSTSUBSCRIPT sec end_POSTSUBSCRIPT. However, as we keep lowering \mathcal{R}caligraphic_R by increasing MBHsubscript𝑀BHM_{\text{BH}}italic_M start_POSTSUBSCRIPT BH end_POSTSUBSCRIPT, binaries start to become unstable very fast according to equation (19) and hence the likelihood of them merging before they evaporate decreases. We also move into a regime in which the binaries only experience a few flyby interactions in each cycle, which could also contribute to decreasing the probability of reaching extreme eccentricity values as occurs in figure 6. Our experiments suggest that there is an optimal value for 0.02similar-to0.02\mathcal{R}\sim 0.02caligraphic_R ∼ 0.02 at which the merger fraction peaks.

5.2 Understanding the optimal value of \mathcal{R}caligraphic_R to increase the merger fractions

We provide an analytical explanation at the order-of-magnitude level of why there is a particular value for \mathcal{R}caligraphic_R at which the merger fraction peaks as shown in figure 6.

We can write our merger condition as jmin=1emax2subscript𝑗min1superscriptsubscript𝑒max2j_{\text{min}}=\sqrt{1-e_{\text{max}}^{2}}italic_j start_POSTSUBSCRIPT min end_POSTSUBSCRIPT = square-root start_ARG 1 - italic_e start_POSTSUBSCRIPT max end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG reaching values below a critical specific angular momentum jcritsubscript𝑗critj_{\text{crit}}italic_j start_POSTSUBSCRIPT crit end_POSTSUBSCRIPT. The latter is set to jcrit0.02similar-to-or-equalssubscript𝑗crit0.02j_{\text{crit}}\simeq 0.02italic_j start_POSTSUBSCRIPT crit end_POSTSUBSCRIPT ≃ 0.02 in our simulations (1ecrit=1041subscript𝑒critsuperscript1041-e_{\rm crit}=10^{-4}1 - italic_e start_POSTSUBSCRIPT roman_crit end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT). In the absence of flybys (=00\mathcal{R}=0caligraphic_R = 0), the fraction of mergers for initially circular and isotropic orbits is simply111The merger fraction is given by the probability that jmin=5/3|cosI|<jcritsubscript𝑗min53𝐼subscript𝑗critj_{\text{min}}=\sqrt{5/3}|\cos I|<j_{\rm crit}italic_j start_POSTSUBSCRIPT min end_POSTSUBSCRIPT = square-root start_ARG 5 / 3 end_ARG | roman_cos italic_I | < italic_j start_POSTSUBSCRIPT roman_crit end_POSTSUBSCRIPT, which becomes 3/5jcrit35subscript𝑗crit\sqrt{3/5}j_{\rm crit}square-root start_ARG 3 / 5 end_ARG italic_j start_POSTSUBSCRIPT roman_crit end_POSTSUBSCRIPT for an isotropic distribution. 3/5jcrit0.015similar-to-or-equals35subscript𝑗crit0.015\sqrt{3/5}j_{\rm crit}\simeq 0.015square-root start_ARG 3 / 5 end_ARG italic_j start_POSTSUBSCRIPT roman_crit end_POSTSUBSCRIPT ≃ 0.015 in our simulations. This changes slightly to 0.52jcrit0.01similar-to-or-equalsabsent0.52subscript𝑗critsimilar-to-or-equals0.01\simeq 0.52j_{\rm crit}\simeq 0.01≃ 0.52 italic_j start_POSTSUBSCRIPT roman_crit end_POSTSUBSCRIPT ≃ 0.01 for an initial thermal distribution.

As Figure 4 shows, the dominant effect of flybys on the merger rates comes from changes in inclinations, which corresponds to changes in jzsubscript𝑗𝑧j_{z}italic_j start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT at fixed j𝑗jitalic_j. Binaries that reach |jz|<jcritsubscript𝑗𝑧subscript𝑗crit|j_{z}|<j_{\text{crit}}| italic_j start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT | < italic_j start_POSTSUBSCRIPT crit end_POSTSUBSCRIPT can merge, provided that the binary lies in that range for at least a significant fraction of a secular timescale so it can be assisted by the tidal field. In a completely diffusive scenario, jzsubscript𝑗𝑧j_{z}italic_j start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT undergoes a random walk due to stellar encounters such that jz2=𝒟tdelimited-⟨⟩superscriptsubscript𝑗𝑧2𝒟𝑡\langle j_{z}^{2}\rangle=\mathcal{D}t⟨ italic_j start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ = caligraphic_D italic_t, with 𝒟𝒟\mathcal{D}caligraphic_D a constant diffusive parameter. Thus, in a relaxation timescale 1/𝒟1𝒟1/\mathcal{D}1 / caligraphic_D binaries outside the range |jz|<jcritsubscript𝑗𝑧subscript𝑗crit|j_{z}|<j_{\text{crit}}| italic_j start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT | < italic_j start_POSTSUBSCRIPT crit end_POSTSUBSCRIPT will be kicked in, monotonically increasing the fraction of mergers with 𝒟proportional-to𝒟\mathcal{R}\propto\mathcal{D}caligraphic_R ∝ caligraphic_D consistent with the increase in Figure 6 below 0.02similar-to0.02\mathcal{R}\sim 0.02caligraphic_R ∼ 0.02.

As we continue to increase the diffusion rate due to stellar encounters 𝒟𝒟\mathcal{D}caligraphic_D, the merger fraction is expected to turn over when the binaries are kicked in and out of the |jz|<jcritsubscript𝑗𝑧subscript𝑗crit|j_{z}|<j_{\text{crit}}| italic_j start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT | < italic_j start_POSTSUBSCRIPT crit end_POSTSUBSCRIPT range in timescale shorter than the secular time. The critical diffusion timescale tcritsubscript𝑡critt_{\text{crit}}italic_t start_POSTSUBSCRIPT crit end_POSTSUBSCRIPT to linger in this region is given by

tcritjcrit2/𝒟.similar-tosubscript𝑡critsuperscriptsubscript𝑗crit2𝒟t_{\text{crit}}\sim j_{\text{crit}}^{2}/\mathcal{D}.italic_t start_POSTSUBSCRIPT crit end_POSTSUBSCRIPT ∼ italic_j start_POSTSUBSCRIPT crit end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / caligraphic_D . (23)

Thus, we expect flybys to most efficiently boost the merger rate when tcritτ0similar-tosubscript𝑡critsubscript𝜏0t_{\text{crit}}\sim\tau_{0}italic_t start_POSTSUBSCRIPT crit end_POSTSUBSCRIPT ∼ italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, a condition that can be expressed in terms of our definition of \mathcal{R}caligraphic_R as

=(τ0𝒟)1/2jcrit,superscriptsubscript𝜏0𝒟12similar-tosubscript𝑗crit\mathcal{R}=(\tau_{0}\mathcal{D})^{1/2}\sim j_{\text{crit}},caligraphic_R = ( italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT caligraphic_D ) start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT ∼ italic_j start_POSTSUBSCRIPT crit end_POSTSUBSCRIPT , (24)

which would be the value of \mathcal{R}caligraphic_R for which the merger fraction peaks. In our simulations, this estimate predicts a peak value at 0.02similar-to0.02\mathcal{R}\sim 0.02caligraphic_R ∼ 0.02, which perfectly matches the observed peak in Figure 6.

However, it is important to note that our simulations are not in a fully diffusive regime and we observe sudden large jumps in ainsubscript𝑎ina_{\text{in}}italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT (see Figure 5) due to penetrative encounters (b<ain𝑏subscript𝑎inb<a_{\text{in}}italic_b < italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT). Therefore, Equation (24) serves as a general approximation to where the merger fraction should peak, but can be slightly off due to the non-diffusive nature of these penetrative flybys. In order to illustrate the approximate nature of our estimates, we repeat the numerical experiments for different values of jcritsubscript𝑗critj_{\rm crit}italic_j start_POSTSUBSCRIPT roman_crit end_POSTSUBSCRIPT in Figure B1. We observe that even though there is a value for \mathcal{R}caligraphic_R at which the merger fraction peaks, and it decreases for lower values of jcritsubscript𝑗critj_{\text{crit}}italic_j start_POSTSUBSCRIPT crit end_POSTSUBSCRIPT, the peaks do not occur exactly at jcritsimilar-tosubscript𝑗crit\mathcal{R}\sim j_{\text{crit}}caligraphic_R ∼ italic_j start_POSTSUBSCRIPT crit end_POSTSUBSCRIPT.

6 Lifting the relativistic quenching from flybys

Considering the properties of the GC, it is likely that a considerable amount of binaries residing in it will be subject to relativistic quenching. This presents an obstacle when producing mergers, as it does not allow binaries to reach extreme eccentricities. We can quantify the relative importance of relativistic precession using the following dimensionless parameter ϵGRsubscriptitalic-ϵGR\epsilon_{\text{GR}}italic_ϵ start_POSTSUBSCRIPT GR end_POSTSUBSCRIPT as (Petrovich & Antonini, 2017):

ϵGRτsecτGR=4GMbin2aout3(1eout2)3/2c2ain4MBH×[1+MclMBH2aout3γ(aout+s)3γ(23aout+sγ2(aout+s))]1,subscriptitalic-ϵGRsubscript𝜏secsubscript𝜏GR4𝐺superscriptsubscript𝑀bin2superscriptsubscript𝑎out3superscript1superscriptsubscript𝑒out232superscript𝑐2superscriptsubscript𝑎in4subscript𝑀BHsuperscriptdelimited-[]1subscript𝑀clsubscript𝑀BH2superscriptsubscript𝑎out3𝛾superscriptsubscript𝑎out𝑠3𝛾23subscript𝑎out𝑠𝛾2subscript𝑎out𝑠1\begin{split}\epsilon_{\text{GR}}&\equiv\frac{\tau_{\text{sec}}}{\tau_{\text{% GR}}}\\ &=\frac{4GM_{\text{bin}}^{2}a_{\text{out}}^{3}(1-e_{\text{out}}^{2})^{3/2}}{c^% {2}a_{\text{in}}^{4}M_{\text{BH}}}\\ &\times\left[1+\frac{M_{\text{cl}}}{M_{\text{BH}}}\frac{2a_{\text{out}}^{3-% \gamma}}{(a_{\text{out}}+s)^{3-\gamma}}\left(2-\frac{3a_{\text{out}}+s\gamma}{% 2(a_{\text{out}}+s)}\right)\right]^{-1},\end{split}start_ROW start_CELL italic_ϵ start_POSTSUBSCRIPT GR end_POSTSUBSCRIPT end_CELL start_CELL ≡ divide start_ARG italic_τ start_POSTSUBSCRIPT sec end_POSTSUBSCRIPT end_ARG start_ARG italic_τ start_POSTSUBSCRIPT GR end_POSTSUBSCRIPT end_ARG end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL = divide start_ARG 4 italic_G italic_M start_POSTSUBSCRIPT bin end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ( 1 - italic_e start_POSTSUBSCRIPT out end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 3 / 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_c start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT BH end_POSTSUBSCRIPT end_ARG end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL × [ 1 + divide start_ARG italic_M start_POSTSUBSCRIPT cl end_POSTSUBSCRIPT end_ARG start_ARG italic_M start_POSTSUBSCRIPT BH end_POSTSUBSCRIPT end_ARG divide start_ARG 2 italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 - italic_γ end_POSTSUPERSCRIPT end_ARG start_ARG ( italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT + italic_s ) start_POSTSUPERSCRIPT 3 - italic_γ end_POSTSUPERSCRIPT end_ARG ( 2 - divide start_ARG 3 italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT + italic_s italic_γ end_ARG start_ARG 2 ( italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT + italic_s ) end_ARG ) ] start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT , end_CELL end_ROW (25)

where binaries with values of ϵGR>0.1subscriptitalic-ϵGR0.1\epsilon_{\text{GR}}>0.1italic_ϵ start_POSTSUBSCRIPT GR end_POSTSUBSCRIPT > 0.1 are considered as quenched.

When this parameter becomes smaller relativistic quenching can be lifted, allowing for a potential merger driven by the tidal fields. This can be achieved through the softening of binaries, as this will increase their secular timescale τsecsubscript𝜏sec\tau_{\rm sec}italic_τ start_POSTSUBSCRIPT roman_sec end_POSTSUBSCRIPT. Conveniently, this paper has focused on the effect of flybys on binaries in the GC, which tend to soften said binaries and hence can effectively lift the effects of relativistic quenching.

In figure 7 we can see a specific example where softening driven by flybys lifts relativistic quenching and allows for a merger. The initial conditions of the simulated binary are ain=1subscript𝑎in1a_{\text{in}}=1italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT = 1au and aout=0.3subscript𝑎out0.3a_{\text{out}}=0.3italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT = 0.3pc, which makes it a soft binary severely quenched by GR with an initial dimensionless parameter ϵGR200similar-tosubscriptitalic-ϵGR200\epsilon_{\text{GR}}\sim 200italic_ϵ start_POSTSUBSCRIPT GR end_POSTSUBSCRIPT ∼ 200 . However, due to flyby interactions increasing ainsubscript𝑎ina_{\text{in}}italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT, ϵGRsubscriptitalic-ϵGR\epsilon_{\text{GR}}italic_ϵ start_POSTSUBSCRIPT GR end_POSTSUBSCRIPT reaches values smaller than 0.1 hence lifting relativistic quenching. It is clear that while subject to GR quenching the eccentricity of the binary’s orbit remains practically constant at a low value, such that ainqinsimilar-tosubscript𝑎insubscript𝑞ina_{\text{in}}\sim q_{\text{in}}italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT ∼ italic_q start_POSTSUBSCRIPT in end_POSTSUBSCRIPT. However once it escapes the GR quenching regime, extreme eccentricity oscillations are induced such that the pericentric distance reaches values as low as 103superscript10310^{-3}10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPTau, allowing for a potential merger.

7 Discussion

We have shown that stellar flybys can greatly enhance the merger rates of binaries in the centers of galaxies by assisting the eccentricity growth to extreme values that are driven by the tidal torques exerted by the SMBH and the central cluster.

More specifically, our results demonstrate that the diffusive effects of flybys and the torquing of binary’s orbits due to tidal fields produce comparable merger rates (within a factor of 3; see Figure 3). In turn, for regimes in which the diffusion rate is much slower than the tidal field torquing rate (0.01similar-to0.01\mathcal{R}\sim 0.01caligraphic_R ∼ 0.01; see §5) flybys added to the tidal fields are able to produce a merger rate 30similar-toabsent30\sim 30∼ 30 times larger that the one produced by tidal fields alone. This is due to the chaotic behavior introduced by flyby interactions, as they constantly reset the binary’s orbital parameters and consequently can produce random and dramatic changes in the orbit. Most importantly, the diffusion of the zlimit-from𝑧z-italic_z -component of the angular momentum jzsubscript𝑗𝑧j_{z}italic_j start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT.

We also find that the softening of GR quenched binaries due to flyby interactions allows them to escape the relativistic quenching regime, thus expanding the available phase-space for mergers.

7.1 Dynamical regimes in the galactic center

In Figure 8 we show the different dynamical regimes as a function of ainsubscript𝑎ina_{\text{in}}italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT and aoutsubscript𝑎outa_{\text{out}}italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT, along with a contour plot of \mathcal{R}caligraphic_R. It is important to note that in this plot, Equation (22) has been adapted such that vpsubscript𝑣𝑝v_{p}italic_v start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT, nsubscript𝑛n_{\star}italic_n start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT and f(β)𝑓𝛽f(\beta)italic_f ( italic_β ) are written in terms of aoutsubscript𝑎outa_{\text{out}}italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT. We can see that for binaries that are extremely tight (ain0.1less-than-or-similar-tosubscript𝑎in0.1a_{\text{in}}\lesssim 0.1italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT ≲ 0.1au), \mathcal{R}caligraphic_R reaches very low values which implies an almost insignificant diffusion in the angular momentum. If these binaries are subject to relativistic quenching, even if they are soft, it is unlikely that flyby interactions will have enough impact on their orbit to push them out of this regime. Considering this, the lifting of GR quenching due to flyby interactions is probably most efficient when ain0.1greater-than-or-equivalent-tosubscript𝑎in0.1a_{\text{in}}\gtrsim 0.1italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT ≳ 0.1 au

Our results suggest that previous studies such as Petrovich & Antonini (2017) and Hoang et al. (2018) (who considered a regime similar to our own with aout0.5less-than-or-similar-tosubscript𝑎out0.5a_{\text{out}}\lesssim 0.5italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT ≲ 0.5pc and ϵGR0.1less-than-or-similar-tosubscriptitalic-ϵGR0.1\epsilon_{\text{GR}}\lesssim 0.1italic_ϵ start_POSTSUBSCRIPT GR end_POSTSUBSCRIPT ≲ 0.1) have underestimated the effect of interactions with field stars, presenting it as merely a limiting evaporation timescale that should inhibit merger rates. Our findings on the other hand, suggest that flyby interactions are actually quite efficient when producing mergers, much more so than tidal fields alone. This is due mainly to the diffusion in the angular momentum vector, quantified by our dimensionless parameter \mathcal{R}caligraphic_R, that for these previous works would be of order 0.01similar-to0.01\mathcal{R}\sim 0.01caligraphic_R ∼ 0.01. This, combined with the ability to lift relativistic quenching and hence increase the phase-space for mergers (which was not taken into account in the mentioned studies either), provides strong evidence for the importance of flybys in producing mergers in the GC and shows that their effects must be accounted for.

7.2 Other astrophysical environments

Our results are applicable the evolution of wide binaries in various astrophysical environments. Next we discuss our results in the context of previous results.

Solar neighborhood

The dynamical evolution of wide binaries is equivalent to that of comets in the Oort Cloud studied by Heisler & Tremaine (1986). This is a different regime than the one considered in this paper, where the evolution should be flyby dominated (higher value for \mathcal{R}caligraphic_R) due to the orbits of binaries being wider and having longer secular timescales. We can write an expression for \mathcal{R}caligraphic_R in this environment as follows

9(MbinM)14(Mp0.162M)(vp40km/s)12×(ain2.5×104au)34(n1.14pc3)12(ρ00.185Mpc3)12,similar-to-or-equals9superscriptsubscript𝑀binsubscript𝑀direct-product14subscript𝑀𝑝0.162subscript𝑀direct-productsuperscriptsubscript𝑣𝑝40km/s12superscriptsubscript𝑎in2.5superscript104au34superscriptsubscript𝑛1.14superscriptpc312superscriptsubscript𝜌00.185subscript𝑀direct-productsuperscriptpc312\begin{split}\mathcal{R}\simeq&9\left(\frac{M_{\text{bin}}}{M_{\odot}}\right)^% {-\frac{1}{4}}\left(\frac{M_{p}}{0.162M_{\odot}}\right)\left(\frac{v_{p}}{40% \text{km/s}}\right)^{-\frac{1}{2}}\\ &\times\left(\frac{a_{\text{in}}}{2.5\times 10^{4}\text{au}}\right)^{\frac{3}{% 4}}\left(\frac{n_{\star}}{1.14\text{pc}^{-3}}\right)^{\frac{1}{2}}\left(\frac{% \rho_{0}}{0.185M_{\odot}\text{pc}^{-3}}\right)^{-\frac{1}{2}},\end{split}start_ROW start_CELL caligraphic_R ≃ end_CELL start_CELL 9 ( divide start_ARG italic_M start_POSTSUBSCRIPT bin end_POSTSUBSCRIPT end_ARG start_ARG italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT - divide start_ARG 1 end_ARG start_ARG 4 end_ARG end_POSTSUPERSCRIPT ( divide start_ARG italic_M start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG start_ARG 0.162 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT end_ARG ) ( divide start_ARG italic_v start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG start_ARG 40 km/s end_ARG ) start_POSTSUPERSCRIPT - divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL × ( divide start_ARG italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT end_ARG start_ARG 2.5 × 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT au end_ARG ) start_POSTSUPERSCRIPT divide start_ARG 3 end_ARG start_ARG 4 end_ARG end_POSTSUPERSCRIPT ( divide start_ARG italic_n start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG 1.14 pc start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT end_ARG ) start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT ( divide start_ARG italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG 0.185 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT pc start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT end_ARG ) start_POSTSUPERSCRIPT - divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT , end_CELL end_ROW (26)

where we have considered the secular timescale as

τg=Mbin1/22πρ0G1/2ain3/2,subscript𝜏𝑔superscriptsubscript𝑀bin122𝜋subscript𝜌0superscript𝐺12superscriptsubscript𝑎in32\tau_{g}=\frac{M_{\text{bin}}^{1/2}}{2\pi\rho_{0}G^{1/2}a_{\text{in}}^{3/2}},italic_τ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT = divide start_ARG italic_M start_POSTSUBSCRIPT bin end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_π italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_G start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 / 2 end_POSTSUPERSCRIPT end_ARG , (27)

with ρ0=0.185±0.02Mpc3subscript𝜌0plus-or-minus0.1850.02subscript𝑀direct-productsuperscriptpc3\rho_{0}=0.185\pm 0.02M_{\odot}\text{pc}^{-3}italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.185 ± 0.02 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT pc start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT being the local density as used by Heisler & Tremaine (1986).

We can see that for an average comet in the Oort Cloud we get 9similar-to-or-equals9\mathcal{R}\simeq 9caligraphic_R ≃ 9 as opposed to the 0.02similar-to-or-equals0.02\mathcal{R}\simeq 0.02caligraphic_R ≃ 0.02 obtained for compact object binaries in the GC. This implies that in the Oort Cloud flybys dominate over tidal fields, as the diffusion rate from these interactions is 9similar-toabsent9\sim 9∼ 9 times faster than the torquing due to the tidal field. This is mainly due to the difference in the amount of flybys in a secular timescale, τ0/tencsubscript𝜏0subscript𝑡enc\tau_{0}/t_{\mathrm{enc}}italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_t start_POSTSUBSCRIPT roman_enc end_POSTSUBSCRIPT. As shown in equation (A7), we get about 68 encounters for an average binary in the GC, while in the solar neighbourhood the average comet suffers 2×104similar-toabsent2superscript104\sim 2\times 10^{4}∼ 2 × 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT flyby encounters (considering equation (27) as τ0subscript𝜏0\tau_{0}italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and tenc0.015Myrsimilar-to-or-equalssubscript𝑡enc0.015Myrt_{\mathrm{enc}}\simeq 0.015\mathrm{Myr}italic_t start_POSTSUBSCRIPT roman_enc end_POSTSUBSCRIPT ≃ 0.015 roman_Myr as used by Heisler & Tremaine 1986). This increase in the amount of close encounters inevitably results in more diffusion of the angular momentum j𝑗\vec{j}over→ start_ARG italic_j end_ARG, hence the higher value of \mathcal{R}caligraphic_R obtained in this regime. Consistently, the numerical experiments by Heisler & Tremaine (1986) show the secular evolution is largely detuned due to the stellar flybys, making it hard to complete a clean secular cycle in the ωe𝜔𝑒\omega-eitalic_ω - italic_e space (see figure 4 therein). This in turn makes it difficult to achieve the high eccentricities required by a merger, as is discussed by Stegmann et al. (2024).

Globular clusters

A recent study by Rasskazov & Rafikov (2023) looked at the evolution of compact object binaries when subject to tidal fields and stellar encounters in globular clusters, similar to the work done throughout this paper. The main difference is that these globular clusters are less dense and do not contain a central SMBH as the NSCs we have considered in this study, which means that they have a much lower velocity dispersion. This results in a higher ainsubscript𝑎ina_{\text{in}}italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT hard-soft limit for binaries, implying that many of the binaries in these environments are actually hard, unlike our work. This is particularly important when speaking of GR quenching, as we have found that flybys are able to diminish this effect only when they are capable of softening binaries. In the case of hard binaries, flybys should tend to promote GR quenching rather than lift it, further inhibiting merger rates.

By looking at the behavior in the various evolution examples in Rasskazov & Rafikov (2023), we suggest that similar to the galactic field, the diffusion parameter \mathcal{R}caligraphic_R is large (dominated by flybys). We note that equation (22) is not applicable here because the encounters are generally far from being impulsive given the low dispersion velocities. More work would need to be done to extend the analysis to these environment, including flybys that can be secular in nature (e.g., Heggie & Rasio 1996; Hamers & Samsing 2019).

8 Conclusions

In this work we have studied the combined effect of cluster tidal fields and flyby interactions on the evolution of binaries in the galactic center. Our main result is the dramatic synergy between these two physical processes at driving binaries into extremely high eccentricities (1emax104less-than-or-similar-to1subscript𝑒maxsuperscript1041-e_{\text{max}}\lesssim 10^{-4}1 - italic_e start_POSTSUBSCRIPT max end_POSTSUBSCRIPT ≲ 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT) and their subsequent merger.

Although stellar flybys tend to increase the semi-major axis of the binaries thus inhibiting the mergers, they also give rise an stochastic evolution of other orbital elements that can dramatically shrink the pericenter distance before the binary becomes unbound. This dynamics has been previously studied showing that it could lead to mergers of wide binaries in the galactic field, consistent with our results for the galactic center including only flybys that drive mergers in few a percent of the systems. We extend these results and show that this rate increases by an order of magnitude when flybys act in concert with the static tidal field from the SMBH and the central cluster.

We find that the main cause of this synergy is the persistent tidal field-driven eccentricity excitation that is enhanced by the diffusion of jzsubscript𝑗𝑧j_{z}italic_j start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT due to flyby encounters. We calculated the diffusion rate of the angular momentum vector j𝑗\vec{j}over→ start_ARG italic_j end_ARG due to flybys using the impulse approximation and assess the various regimes with a dimensionless parameter \mathcal{R}caligraphic_R (Eq. 22) that measures the expected diffusion within a secular timesale (i.e, torquing timescale due to tidal fields). Our experiments suggest that the merger rates peak in the range 0.010.1similar-to0.010.1\mathcal{R}\sim 0.01-0.1caligraphic_R ∼ 0.01 - 0.1, meaning when the diffusion rate is 10100similar-toabsent10100\sim 10-100∼ 10 - 100 times more slow than the torquing due to the tidal field. This regime is typical of the central parsec of our GC (see Figure 8), while other environments such the solar neighborhood and stellar clusters tend to have much higher values of \mathcal{R}caligraphic_R.

We also show that the gradual softening of binaries in the GC can lift the relativistic quenching of initially tight binaries, thus expanding the available phase-space for mergers. This is also likely to contribute to the previously discussed synergy, given that by increasing the amount of binaries that can potentially merge in a population we can expect a higher merger rate.

In summary, we conclude that despite the gradual softening of binaries due to stellar encounters, these greatly enhance merger rates in GCs by promoting the tidal field driven eccentricity excitation This behavior has been ignored in previous works and further reinforces that galactic centers are ideal environments for the production of compact object binary mergers.

We would like to thank Antranik Sefilian, Carolina Charalambous, Diego Muñoz, Mark Dodici, and Scott Tremaine for useful discussions. We are also grateful to an anonymous referee for providing helpful comments on the maniscript. CP acknowledges support from CATA-Basal AFB-170002, ANID BASAL project FB210003, FONDECYT Regular grant 1210425, CASSACA grant CCJRF2105, and ANID+REC Convocatoria Nacional subvencion a la instalacion en la Academia convocatoria 2020 PAI77200076. CH is supported by the John N. Bahcall Fellowship Fund at the Institute for Advanced Study.

References

  • Abbott et al. (2016) Abbott, B. P., Abbott, R., Abbott, T. D., et al. 2016, Phys. Rev. Lett., 116, 061102, doi: 10.1103/PhysRevLett.116.061102
  • Abbott et al. (2017) —. 2017, Phys. Rev. Lett., 119, 161101, doi: 10.1103/PhysRevLett.119.161101
  • Alexander & Pfuhl (2013) Alexander, T., & Pfuhl, O. 2013, The Astrophysical Journal, 780, 148, doi: 10.1088/0004-637x/780/2/148
  • Alexander & Pfuhl (2014) Alexander, T., & Pfuhl, O. 2014, ApJ, 780, 148, doi: 10.1088/0004-637X/780/2/148
  • Antonini & Perets (2012) Antonini, F., & Perets, H. B. 2012, The Astrophysical Journal, 757, 27, doi: 10.1088/0004-637x/757/1/27
  • Binney & Tremaine (2008) Binney, J., & Tremaine, S. 2008, Galactic Dynamics: Second Edition
  • Bub & Petrovich (2020) Bub, M. W., & Petrovich, C. 2020, ApJ, 894, 15, doi: 10.3847/1538-4357/ab8461
  • Collins & Sari (2008) Collins, B. F., & Sari, R. 2008, The Astronomical Journal, 136, 2552, doi: 10.1088/0004-6256/136/6/2552
  • Gautam et al. (2024) Gautam, A. K., Do, T., Ghez, A. M., et al. 2024, arXiv e-prints, arXiv:2401.12555, doi: 10.48550/arXiv.2401.12555
  • Georgiev & Böker (2014) Georgiev, I. Y., & Böker, T. 2014, Monthly Notices of the Royal Astronomical Society, 441, 3570, doi: 10.1093/mnras/stu797
  • Ghez et al. (2005) Ghez, A. M., Salim, S., Hornstein, S. D., et al. 2005, The Astrophysical Journal, 620, 744–757, doi: 10.1086/427175
  • Gillessen et al. (2009) Gillessen, S., Eisenhauer, F., Trippe, S., et al. 2009, ApJ, 692, 1075, doi: 10.1088/0004-637X/692/2/1075
  • Grishin et al. (2016) Grishin, E., Perets, H. B., Zenati, Y., & Michaely, E. 2016, Monthly Notices of the Royal Astronomical Society, 466, 276, doi: 10.1093/mnras/stw3096
  • Hamers & Samsing (2019) Hamers, A. S., & Samsing, J. 2019, MNRAS, 487, 5630, doi: 10.1093/mnras/stz1646
  • Hamers & Tremaine (2017) Hamers, A. S., & Tremaine, S. 2017, AJ, 154, 272, doi: 10.3847/1538-3881/aa9926
  • Hamilton & Modak (2023) Hamilton, C., & Modak, S. 2023. https://arxiv.org/abs/2311.04352
  • Hamilton & Rafikov (2019) Hamilton, C., & Rafikov, R. R. 2019, ApJ, 881, L13, doi: 10.3847/2041-8213/ab3468
  • Hamilton & Rafikov (2019a) Hamilton, C., & Rafikov, R. R. 2019a, Monthly Notices of the Royal Astronomical Society, 488, 5489, doi: 10.1093/mnras/stz1730
  • Hamilton & Rafikov (2019b) —. 2019b, Monthly Notices of the Royal Astronomical Society, 488, 5512, doi: 10.1093/mnras/stz2026
  • Hamilton & Rafikov (2021) —. 2021, MNRAS, 505, 4151, doi: 10.1093/mnras/stab1284
  • Hamilton & Rafikov (2023) Hamilton, C., & Rafikov, R. R. 2023, arXiv e-prints, arXiv:2306.03703, doi: 10.48550/arXiv.2306.03703
  • Heggie (1975) Heggie, D. C. 1975, MNRAS, 173, 729, doi: 10.1093/mnras/173.3.729
  • Heggie & Rasio (1996) Heggie, D. C., & Rasio, F. A. 1996, MNRAS, 282, 1064, doi: 10.1093/mnras/282.3.1064
  • Heisler & Tremaine (1986) Heisler, J., & Tremaine, S. 1986, Icarus, 65, 13, doi: https://doi.org/10.1016/0019-1035(86)90060-6
  • Hills (1975a) Hills, J. G. 1975a, AJ, 80, 809, doi: 10.1086/111815
  • Hills (1975b) —. 1975b, AJ, 80, 1075, doi: 10.1086/111842
  • Hoang et al. (2019) Hoang, B.-M., Naoz, S., Kocsis, B., Farr, W. M., & McIver, J. 2019, The Astrophysical Journal, 875, L31, doi: 10.3847/2041-8213/ab14f7
  • Hoang et al. (2018) Hoang, B.-M., Naoz, S., Kocsis, B., Rasio, F. A., & Dosopoulou, F. 2018, ApJ, 856, 140, doi: 10.3847/1538-4357/aaafce
  • Holman et al. (1997) Holman, M., Touma, J., & Tremaine, S. 1997, Nature, 386, 254, doi: 10.1038/386254a0
  • Kocsis & Tremaine (2011) Kocsis, B., & Tremaine, S. 2011, Monthly Notices of the Royal Astronomical Society, 412, 187, doi: 10.1111/j.1365-2966.2010.17897.x
  • Leigh et al. (2017) Leigh, N. W. C., Geller, A. M., McKernan, B., et al. 2017, Monthly Notices of the Royal Astronomical Society, 474, 5672, doi: 10.1093/mnras/stx3134
  • Michaely & Perets (2020) Michaely, E., & Perets, H. B. 2020, MNRAS, 498, 4924, doi: 10.1093/mnras/staa2720
  • Modak & Hamilton (2023) Modak, S., & Hamilton, C. 2023, Monthly Notices of the Royal Astronomical Society, 524, 3102, doi: 10.1093/mnras/stad2073
  • Naoz (2016) Naoz, S. 2016, ARA&A, 54, 441, doi: 10.1146/annurev-astro-081915-023315
  • Naoz et al. (2011) Naoz, S., Farr, W. M., Lithwick, Y., Rasio, F. A., & Teyssandier, J. 2011, Nature, 473, 187, doi: 10.1038/nature10076
  • Naoz et al. (2013) —. 2013, Monthly Notices of the Royal Astronomical Society, 431, 2155, doi: 10.1093/mnras/stt302
  • Neumayer et al. (2011) Neumayer, N., Walcher, C. J., Andersen, D., et al. 2011, MNRAS, 413, 1875, doi: 10.1111/j.1365-2966.2011.18266.x
  • Peters (1964) Peters, P. C. 1964, Phys. Rev., 136, B1224, doi: 10.1103/PhysRev.136.B1224
  • Petrovich & Antonini (2017) Petrovich, C., & Antonini, F. 2017, The Astrophysical Journal, 846, 146, doi: 10.3847/1538-4357/aa8628
  • Rasskazov & Rafikov (2023) Rasskazov, A., & Rafikov, R. R. 2023, arXiv e-prints, arXiv:2310.15374. https://arxiv.org/abs/2310.15374
  • Rein & Liu (2012) Rein, H., & Liu, S.-F. 2012, A&A, 537, A128, doi: 10.1051/0004-6361/201118085
  • Rose et al. (2020) Rose, S. C., Naoz, S., Gautam, A. K., et al. 2020, The Astrophysical Journal, 904, 113, doi: 10.3847/1538-4357/abc557
  • Schödel et al. (2014) Schödel, R., Feldmeier, A., Kunneriath, D., et al. 2014, A&A, 566, A47, doi: 10.1051/0004-6361/201423481
  • Schödel et al. (2018) Schödel, R., Gallego-Cano, E., Dong, H., et al. 2018, A&A, 609, A27, doi: 10.1051/0004-6361/201730452
  • Stegmann et al. (2024) Stegmann, J., Vigna-Gómez, A., Rantala, A., et al. 2024, Close Encounters of Wide Binaries Induced by the Galactic Tide: Implications for Stellar Mergers and Gravitational-Wave Sources. https://arxiv.org/abs/2405.02912
  • Stephan et al. (2016) Stephan, A. P., Naoz, S., Ghez, A. M., et al. 2016, MNRAS, 460, 3494, doi: 10.1093/mnras/stw1220
  • Stephan et al. (2019) Stephan, A. P., Naoz, S., Ghez, A. M., et al. 2019, The Astrophysical Journal, 878, 58, doi: 10.3847/1538-4357/ab1e4d
  • Tremaine (2023) Tremaine, S. 2023, Dynamics of Planetary Systems
  • Turner et al. (2012) Turner, M. L., Côté, P., Ferrarese, L., et al. 2012, ApJS, 203, 5, doi: 10.1088/0067-0049/203/1/5

Appendix A Impulse approximation

From Collins & Sari (2008), we can write the variation in the velocity (impulse) ΔviΔsubscript𝑣𝑖\Delta\vec{v_{i}}roman_Δ over→ start_ARG italic_v start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG of one component of the binary with mass misubscript𝑚𝑖m_{i}italic_m start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT (i={1,2}𝑖12i=\{1,2\}italic_i = { 1 , 2 }) considering an impact parameter bisubscript𝑏𝑖\vec{b_{i}}over→ start_ARG italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG as follows:

Δvi=2Gmpvpbibi^.Δsubscript𝑣𝑖2𝐺subscript𝑚𝑝subscript𝑣𝑝subscript𝑏𝑖^subscript𝑏𝑖\Delta\vec{v_{i}}=\frac{2Gm_{p}}{v_{p}b_{i}}\hat{b_{i}}.roman_Δ over→ start_ARG italic_v start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG = divide start_ARG 2 italic_G italic_m start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG start_ARG italic_v start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG over^ start_ARG italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG . (A1)

The impact parameter and the velocity of the perturber must be orthogonal such that bivp=0subscript𝑏𝑖subscript𝑣𝑝0\vec{b}_{i}\cdot\vec{v}_{p}=0over→ start_ARG italic_b end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⋅ over→ start_ARG italic_v end_ARG start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = 0. The trajectory of the perturber is approximated to a straight line, such that r(t)=bi+vpt𝑟𝑡subscript𝑏𝑖subscript𝑣𝑝𝑡\vec{r}(t)=\vec{b}_{i}+\vec{v}_{p}tover→ start_ARG italic_r end_ARG ( italic_t ) = over→ start_ARG italic_b end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + over→ start_ARG italic_v end_ARG start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT italic_t. The perturber’s velocity is relative to only one component of the binary; therefore, we can relate both impact parameters b1subscript𝑏1b_{1}italic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and b2subscript𝑏2b_{2}italic_b start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT through the equation b2=b1r+v^p(rv^p)subscript𝑏2subscript𝑏1𝑟subscript^𝑣𝑝𝑟subscript^𝑣𝑝\vec{b_{2}}=\vec{b}_{1}-\vec{r}+\hat{v}_{p}(\vec{r}\cdot\hat{v}_{p})over→ start_ARG italic_b start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG = over→ start_ARG italic_b end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - over→ start_ARG italic_r end_ARG + over^ start_ARG italic_v end_ARG start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ( over→ start_ARG italic_r end_ARG ⋅ over^ start_ARG italic_v end_ARG start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ) considering that b2subscript𝑏2\vec{b_{2}}over→ start_ARG italic_b start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG must also be perpendicular to vpsubscript𝑣𝑝\vec{v}_{p}over→ start_ARG italic_v end_ARG start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT. The overall change in relative velocity can then be expressed as Δv=Δv1Δv2Δ𝑣Δsubscript𝑣1Δsubscript𝑣2\Delta\vec{v}=\Delta\vec{v}_{1}-\Delta\vec{v}_{2}roman_Δ over→ start_ARG italic_v end_ARG = roman_Δ over→ start_ARG italic_v end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - roman_Δ over→ start_ARG italic_v end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, where both Δv1Δsubscript𝑣1\Delta\vec{v}_{1}roman_Δ over→ start_ARG italic_v end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and Δv2Δsubscript𝑣2\Delta\vec{v}_{2}roman_Δ over→ start_ARG italic_v end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT obey the equation A1.

We can define ΔvΔ𝑣\Delta\vec{v}roman_Δ over→ start_ARG italic_v end_ARG as the sum of the velocity components perpendicular and parallel to the instantaneous orbital velocity; Δv=Δv+ΔvΔ𝑣Δsubscript𝑣perpendicular-toΔsubscript𝑣parallel-to\Delta\vec{v}=\Delta v_{\perp}+\Delta v_{\parallel}roman_Δ over→ start_ARG italic_v end_ARG = roman_Δ italic_v start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT + roman_Δ italic_v start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT. We can rewrite it using equation A1 as

Δv=2GMpvp(b^1b1b^2b2)Δ𝑣2𝐺subscript𝑀𝑝subscript𝑣𝑝subscript^𝑏1subscript𝑏1subscript^𝑏2subscript𝑏2\Delta\vec{v}=\frac{2GM_{p}}{v_{p}}\left(\frac{\hat{b}_{1}}{b_{1}}-\frac{\hat{% b}_{2}}{b_{2}}\right)roman_Δ over→ start_ARG italic_v end_ARG = divide start_ARG 2 italic_G italic_M start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG start_ARG italic_v start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG ( divide start_ARG over^ start_ARG italic_b end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG italic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG - divide start_ARG over^ start_ARG italic_b end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG start_ARG italic_b start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG ) (A2)
Refer to caption
Figure A1: Geometry used to calculate r^×b^1^𝑟subscript^𝑏1\hat{r}\times\hat{b}_{1}over^ start_ARG italic_r end_ARG × over^ start_ARG italic_b end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and r^×b^2^𝑟subscript^𝑏2\hat{r}\times\hat{b}_{2}over^ start_ARG italic_r end_ARG × over^ start_ARG italic_b end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT. The orbit lies on the equatorial plane of a sphere for reference. The flyby trajectory is approximated to a straight line, and the gray crosses indicate the points where it intersects either the sphere or its equator.

When the perturbing potential is axisymmetric (as is the case in this scenario), the inner angular momentum and eccentricity vectors j𝑗\vec{j}over→ start_ARG italic_j end_ARG and e𝑒\vec{e}over→ start_ARG italic_e end_ARG must always be orthogonal such that je=0𝑗𝑒0\vec{j}\cdot\vec{e}=0over→ start_ARG italic_j end_ARG ⋅ over→ start_ARG italic_e end_ARG = 0. Therefore, if there is a variation in the eccentricity vector due to fly-by encounters, there must also be a variation in j𝑗\vec{j}over→ start_ARG italic_j end_ARG. We can write the change in the angular momentum vector as

Δj=r×ΔvGMbinainΔ𝑗𝑟Δ𝑣𝐺subscript𝑀binsubscript𝑎in\Delta\vec{j}=\frac{\vec{r}\times\Delta\vec{v}}{\sqrt{GM_{\text{bin}}a_{\text{% in}}}}roman_Δ over→ start_ARG italic_j end_ARG = divide start_ARG over→ start_ARG italic_r end_ARG × roman_Δ over→ start_ARG italic_v end_ARG end_ARG start_ARG square-root start_ARG italic_G italic_M start_POSTSUBSCRIPT bin end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT end_ARG end_ARG (A3)

which, using A1, results in

Δj=GMbinain2Mpvp(r×b^1b1r×b^2b2).Δ𝑗𝐺subscript𝑀binsubscript𝑎in2subscript𝑀𝑝subscript𝑣𝑝𝑟subscript^𝑏1subscript𝑏1𝑟subscript^𝑏2subscript𝑏2\Delta\vec{j}=\sqrt{\frac{G}{M_{\text{bin}}a_{\text{in}}}}\frac{2M_{p}}{v_{p}}% \left(\frac{\vec{r}\times\hat{b}_{1}}{b_{1}}-\frac{\vec{r}\times\hat{b}_{2}}{b% _{2}}\right).roman_Δ over→ start_ARG italic_j end_ARG = square-root start_ARG divide start_ARG italic_G end_ARG start_ARG italic_M start_POSTSUBSCRIPT bin end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT end_ARG end_ARG divide start_ARG 2 italic_M start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG start_ARG italic_v start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG ( divide start_ARG over→ start_ARG italic_r end_ARG × over^ start_ARG italic_b end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG italic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG - divide start_ARG over→ start_ARG italic_r end_ARG × over^ start_ARG italic_b end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG start_ARG italic_b start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG ) . (A4)

A.1 Diffusion coefficient analysis

It would be of interest to analyse the relative contributions of the tidal and the impulse elements on the evolution of the angular momentum of the inner binary, for it is an indicator of how likely the binary is to merge at a given moment. To this purpose, we define a diffusion coefficient 𝒟Δj2(f,b,i,ω,Ω)/tenc𝒟subscriptdelimited-⟨⟩superscriptnormΔ𝑗2𝑓𝑏𝑖𝜔Ωsubscript𝑡enc\mathcal{D}\equiv\langle\|\Delta j\|^{2}\rangle_{(f,b,i,\omega,\Omega)}/t_{% \text{enc}}caligraphic_D ≡ ⟨ ∥ roman_Δ italic_j ∥ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT ( italic_f , italic_b , italic_i , italic_ω , roman_Ω ) end_POSTSUBSCRIPT / italic_t start_POSTSUBSCRIPT enc end_POSTSUBSCRIPT. From equation A4, we can write ΔjnormΔ𝑗\|\Delta\vec{j}\|∥ roman_Δ over→ start_ARG italic_j end_ARG ∥ as

Δj=GMbinain2Mpvprr^×b^1b1r^×b^2b2,normΔ𝑗𝐺subscript𝑀binsubscript𝑎in2subscript𝑀𝑝subscript𝑣𝑝𝑟norm^𝑟subscript^𝑏1subscript𝑏1^𝑟subscript^𝑏2subscript𝑏2\|\Delta\vec{j}\|=\sqrt{\frac{G}{M_{\text{bin}}a_{\text{in}}}}\frac{2M_{p}}{v_% {p}}r\left\|\frac{\hat{r}\times\hat{b}_{1}}{b_{1}}-\frac{\hat{r}\times\hat{b}_% {2}}{b_{2}}\right\|,∥ roman_Δ over→ start_ARG italic_j end_ARG ∥ = square-root start_ARG divide start_ARG italic_G end_ARG start_ARG italic_M start_POSTSUBSCRIPT bin end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT end_ARG end_ARG divide start_ARG 2 italic_M start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG start_ARG italic_v start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG italic_r ∥ divide start_ARG over^ start_ARG italic_r end_ARG × over^ start_ARG italic_b end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG italic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG - divide start_ARG over^ start_ARG italic_r end_ARG × over^ start_ARG italic_b end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG start_ARG italic_b start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG ∥ , (A5)

where the term involving the cross products can be evaluated numerically by averaging over 106superscript10610^{6}10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT different values calculated using random impact parameters bisubscript𝑏𝑖b_{i}italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and orbital parameters f𝑓fitalic_f, ω𝜔\omegaitalic_ω, i𝑖iitalic_i and ΩΩ\Omegaroman_Ω so as to cover all possible stages of the binary’s orbit. Using this numerical average, we can write Δj2(f,b,i,ω,Ω)subscriptdelimited-⟨⟩superscriptnormΔ𝑗2𝑓𝑏𝑖𝜔Ω\langle\|\Delta j\|^{2}\rangle_{(f,b,i,\omega,\Omega)}⟨ ∥ roman_Δ italic_j ∥ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT ( italic_f , italic_b , italic_i , italic_ω , roman_Ω ) end_POSTSUBSCRIPT as

Δj2(f,b,i,ω,Ω)6.33×106(Mbin10M)1(MpM)2(vp200km/s)2(ain25au),subscriptdelimited-⟨⟩superscriptdelimited-∥∥Δ𝑗2𝑓𝑏𝑖𝜔Ω6.33superscript106superscriptsubscript𝑀bin10subscript𝑀direct-product1superscriptsubscript𝑀𝑝subscript𝑀direct-product2superscriptsubscript𝑣𝑝200km/s2subscript𝑎in25au\begin{split}\langle\|\Delta\vec{j}\|^{2}\rangle_{(f,b,i,\omega,\Omega)}% \approx&6.33\times 10^{-6}\left(\frac{M_{\text{bin}}}{10M_{\odot}}\right)^{-1}% \left(\frac{M_{p}}{M_{\odot}}\right)^{2}\\ &\left(\frac{v_{p}}{200\text{km/s}}\right)^{-2}\left(\frac{a_{\text{in}}}{25% \text{au}}\right),\end{split}start_ROW start_CELL ⟨ ∥ roman_Δ over→ start_ARG italic_j end_ARG ∥ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT ( italic_f , italic_b , italic_i , italic_ω , roman_Ω ) end_POSTSUBSCRIPT ≈ end_CELL start_CELL 6.33 × 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT ( divide start_ARG italic_M start_POSTSUBSCRIPT bin end_POSTSUBSCRIPT end_ARG start_ARG 10 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( divide start_ARG italic_M start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG start_ARG italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL ( divide start_ARG italic_v start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG start_ARG 200 km/s end_ARG ) start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT ( divide start_ARG italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT end_ARG start_ARG 25 au end_ARG ) , end_CELL end_ROW (A6)

where we have used the average r2=ain2(1+32e2)=74ain2delimited-⟨⟩superscript𝑟2superscriptsubscript𝑎in2132superscript𝑒274superscriptsubscript𝑎in2\langle r^{2}\rangle=a_{\text{in}}^{2}(1+\frac{3}{2}e^{2})=\frac{7}{4}a_{\text% {in}}^{2}⟨ italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ = italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( 1 + divide start_ARG 3 end_ARG start_ARG 2 end_ARG italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) = divide start_ARG 7 end_ARG start_ARG 4 end_ARG italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, considering that e2=0.5delimited-⟨⟩superscript𝑒20.5\langle e^{2}\rangle=0.5⟨ italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ = 0.5 due to the initial thermal distribution assumed.

A.2 Number of flybys

The number of interactions in a secular timescale can be calculated as τ0/tencsubscript𝜏0subscript𝑡enc\tau_{0}/t_{\text{enc}}italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_t start_POSTSUBSCRIPT enc end_POSTSUBSCRIPT, following equations (4) and (17)

τ0tenc67.5(Mbin10M)12(MBH4×106M)1×(aout0.3pc)3(ain25au)12(n106pc3)×(vp200km/s)(f(β)1.8),subscript𝜏0subscript𝑡enc67.5superscriptsubscript𝑀bin10subscript𝑀direct-product12superscriptsubscript𝑀BH4superscript106subscript𝑀direct-product1superscriptsubscript𝑎out0.3pc3superscriptsubscript𝑎in25au12subscript𝑛superscript106superscriptpc3subscript𝑣𝑝200km/s𝑓𝛽1.8\begin{split}\frac{\tau_{0}}{t_{\text{enc}}}&\approx 67.5\left(\frac{M_{\text{% bin}}}{10M_{\odot}}\right)^{\frac{1}{2}}\left(\frac{M_{\text{BH}}}{4\times 10^% {6}M_{\odot}}\right)^{-1}\\ &\times\left(\frac{a_{\text{out}}}{0.3\text{pc}}\right)^{3}\left(\frac{a_{% \text{in}}}{25\text{au}}\right)^{\frac{1}{2}}\left(\frac{n_{\star}}{10^{6}% \text{pc}^{-3}}\right)\\ &\times\left(\frac{v_{p}}{200\text{km/s}}\right)\left(\frac{f(\beta)}{1.8}% \right),\end{split}start_ROW start_CELL divide start_ARG italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG italic_t start_POSTSUBSCRIPT enc end_POSTSUBSCRIPT end_ARG end_CELL start_CELL ≈ 67.5 ( divide start_ARG italic_M start_POSTSUBSCRIPT bin end_POSTSUBSCRIPT end_ARG start_ARG 10 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT ( divide start_ARG italic_M start_POSTSUBSCRIPT BH end_POSTSUBSCRIPT end_ARG start_ARG 4 × 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL × ( divide start_ARG italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT end_ARG start_ARG 0.3 pc end_ARG ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ( divide start_ARG italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT end_ARG start_ARG 25 au end_ARG ) start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT ( divide start_ARG italic_n start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT pc start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT end_ARG ) end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL × ( divide start_ARG italic_v start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG start_ARG 200 km/s end_ARG ) ( divide start_ARG italic_f ( italic_β ) end_ARG start_ARG 1.8 end_ARG ) , end_CELL end_ROW (A7)

where β=s/aout𝛽𝑠subscript𝑎out\beta=s/a_{\text{out}}italic_β = italic_s / italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT, and

f(β)=[12+4(1+3β)(1+β)3]1.𝑓𝛽superscriptdelimited-[]12413𝛽superscript1𝛽31f(\beta)=\left[\frac{1}{2}+\frac{4(1+3\beta)}{(1+\beta)^{3}}\right]^{-1}.italic_f ( italic_β ) = [ divide start_ARG 1 end_ARG start_ARG 2 end_ARG + divide start_ARG 4 ( 1 + 3 italic_β ) end_ARG start_ARG ( 1 + italic_β ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG ] start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT . (A8)

A.3 Numerical experiments for \mathcal{R}caligraphic_R

In order to confirm that the expression derived for \mathcal{R}caligraphic_R using the impulse approximation is accurate, we simulated a binary that evolves only by flyby interactions using REBOUND. From this numerical experiment, in which we considered Mbin=10Msubscript𝑀bin10subscript𝑀direct-productM_{\text{bin}}=10M_{\odot}italic_M start_POSTSUBSCRIPT bin end_POSTSUBSCRIPT = 10 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, Mp=Msubscript𝑀𝑝subscript𝑀direct-productM_{p}=M_{\odot}italic_M start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, MBH=4×106Msubscript𝑀BH4superscript106subscript𝑀direct-productM_{\text{BH}}=4\times 10^{6}M_{\odot}italic_M start_POSTSUBSCRIPT BH end_POSTSUBSCRIPT = 4 × 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, ain=25subscript𝑎in25a_{\text{in}}=25italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT = 25au, aout=0.3subscript𝑎out0.3a_{\text{out}}=0.3italic_a start_POSTSUBSCRIPT out end_POSTSUBSCRIPT = 0.3pc and vp=200subscript𝑣𝑝200v_{p}=200italic_v start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = 200kms-1, we calculated the mean displacement of jnorm𝑗\|\vec{j}\|∥ over→ start_ARG italic_j end_ARG ∥ in different time windows until finding a value at which it converges after one secular timescale τ0subscript𝜏0\tau_{0}italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. This is shown in Figure A2, where we can clearly see how \mathcal{R}caligraphic_R converges to 0.02absent0.02\approx 0.02≈ 0.02 after one secular timescale, which matches the value we retrieve using equation (22) when considering the same initial conditions. In order to keep this simulation as similar as possible to the derivations using the impulse approximation, we do not consider the changes in ainsubscript𝑎ina_{\text{in}}italic_a start_POSTSUBSCRIPT in end_POSTSUBSCRIPT after each encounter, as this would change the amount of flybys per secular timescale.

Refer to caption
Figure A2: Mean square displacement of Δj2delimited-⟨⟩superscriptnormΔ𝑗2\sqrt{\langle\|\Delta\vec{j}\|^{2}\rangle}square-root start_ARG ⟨ ∥ roman_Δ over→ start_ARG italic_j end_ARG ∥ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ end_ARG as a function of time window width w/τ0𝑤subscript𝜏0w/\tau_{0}italic_w / italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT.

Appendix B Merger fraction and jcritsubscript𝑗critj_{\text{crit}}italic_j start_POSTSUBSCRIPT crit end_POSTSUBSCRIPT

To show that the optimal value of \mathcal{R}caligraphic_R at which the merger fraction peaks is proportional to the jcritsubscript𝑗critj_{\text{crit}}italic_j start_POSTSUBSCRIPT crit end_POSTSUBSCRIPT chosen to classify mergers, a set of populations was simulated at different values for \mathcal{R}caligraphic_R. Their merger fractions for three different values of jcritsubscript𝑗critj_{\text{crit}}italic_j start_POSTSUBSCRIPT crit end_POSTSUBSCRIPT are plotted in figure B1 as a function of \mathcal{R}caligraphic_R, and we can see that for smaller jcritsubscript𝑗critj_{\text{crit}}italic_j start_POSTSUBSCRIPT crit end_POSTSUBSCRIPT the peak value tends towards a smaller \mathcal{R}caligraphic_R as well.

Refer to caption
Figure B1: Merger fraction as defined by different values of jcritsubscript𝑗critj_{\text{crit}}italic_j start_POSTSUBSCRIPT crit end_POSTSUBSCRIPT as a function of \mathcal{R}caligraphic_R.