Dynamics of Binary Planets within Star Clusters

Yukun Huang (黄宇坤) National Astronomical Observatory of Japan, 2-21-1 Osawa, Mitaka, Tokyo 181-8588, Japan Department of Astronomy, Tsinghua University, Beijing 100084, China Wei Zhu (祝伟) Department of Astronomy, Tsinghua University, Beijing 100084, China Eiichiro Kokubo (小久保英一郎) National Astronomical Observatory of Japan, 2-21-1 Osawa, Mitaka, Tokyo 181-8588, Japan
Abstract

We develop analytical tools and perform three-body simulations to investigate the orbital evolution and dynamical stability of binary planets within star clusters. Our analytical results show that the orbital stability of a planetary-mass binary against passing stars is mainly related to its orbital period. We find that critical flybys, defined as stellar encounters with energy kicks comparable to the binary binding energy, can efficiently produce a wide range of semimajor axes (a𝑎aitalic_a) and eccentricities (e𝑒eitalic_e) from a dominant population of primordially tight JuMBOs. Applying our results to the recently discovered Jupiter-Mass Binary Objects (JuMBOs) by the James Webb Space Telescope (JWST), our simulations suggest that to match the observed similar-to\sim9% wide binary fraction, an initial semimajor axis of a0=10subscript𝑎010a_{0}=10italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 10–20 au and a density-weighted residence time of χ104greater-than-or-equivalent-to𝜒superscript104\chi\gtrsim 10^{4}italic_χ ≳ 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT Myr pc-3 are favored. These results imply that the JWST JuMBOs probably formed as tight binaries near the cluster core.

Planetary dynamics(2173) – Star clusters(1567) – Planet formation(1241) – Free floating planets(549)
{CJK*}

UTF8gbsn

1 Introduction

With an estimated age of <<<1 Myr and a distance at \approx470 pc (Hillenbrand & Hartmann, 1998), the Trapezium cluster is one of the youngest star clusters near the Sun. It is located at the center of the Orion Nebular Cluster and has one of the densest stellar environments (McCaughrean & Stauffer, 1994). The extreme conditions of the Trapezium Cluster have made it an ideal target for studying the formation of stars and planetary-mass objects (PMOs), both theoretically and observationally (e.g., Prosser et al., 1994; Muench et al., 2008; Pearson & McCaughrean, 2023).

The recent discovery by Pearson & McCaughrean (2023) of 540 PMO candidates, including 40 Jupiter-Mass Binary Objects (JuMBOs), in the Trapezium cluster poses significant challenges to our conventional understanding of planet formation. In the core accretion model, giant planets form within circumstellar protoplanetary disks. The growth from planetesimals to a planetary core (several Msubscript𝑀direct-sumM_{\oplus}italic_M start_POSTSUBSCRIPT ⊕ end_POSTSUBSCRIPT) can last up to a million years, followed by a gas accretion phase lasting up to similar-to\sim10 Myr (Lissauer, 1993; Pollack et al., 1996). This time span is already a factor of a few longer than the estimated age of the Trapezium Cluster. Furthermore, the timescale for ejecting planets out of embedded systems through planet–planet scatterings or stellar encounters is estimated to be another millions to hundreds of millions of years (Spurzem et al., 2009; Nesvorný & Morbidelli, 2012; Huang et al., 2022).

In addition, the abundance of these PMOs relative to stars is also surprisingly high. The Trapezium cluster is estimated to contain similar-to\sim2000 stellar members (Morales-Calderón et al., 2011), so the relative abundance of PMOs is >>>20%percent2020\%20 %. This is above the limit on Jupiter-mass free-floating planets (FFPs) that were derived from microlensing surveys (Mróz et al., 2017). It is also in tension with the substellar mass function derived from Euclid, albeit for a less dense environment (Martín et al., 2024).

What is even more surprising is the discovery of 40 JuMBOs (with binary masses of 2–20 MJsubscript𝑀JM_{\rm J}italic_M start_POSTSUBSCRIPT roman_J end_POSTSUBSCRIPT) at wide separations (d=25𝑑25d=25italic_d = 25–400 au), which contribute similar-to\sim9%percent99\%9 % of all PMOs discovered in the Trapezium Cluster (Pearson & McCaughrean, 2023). Planet-mass binaries have previously been identified. For example, Best et al. (2017) found 2MASS J11193254–1137466AB to be a pair of similar-to\simMJsubscript𝑀𝐽M_{J}italic_M start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT objects with d4𝑑4d\approx 4italic_d ≈ 4 au, and Beichman et al. (2013) found WISE 1828+2650 to be a pair of 3–6 MJsubscript𝑀𝐽M_{J}italic_M start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT objects with d<0.5𝑑0.5d<0.5italic_d < 0.5 au. Oph 162225–240515 is a wide (d242𝑑242d\approx 242italic_d ≈ 242 au) binary that was originally thought to be in the planetary-mass regime (Jayawardhana & Ivanov, 2006), but later studies have substantially revised the masses upward (Luhman et al., 2007).

Despite these previous discoveries, the existence of such abundant (similar-to\sim9%percent99\%9 %) low-mass binaries with wide separations has not been anticipated.

Explaining the formation of binary planets with wide separations is a challenging task. Although two planets (on mutually close astrocentric orbits) can in principle be ejected as a binary pair through stellar encounters (Wang et al., 2024), recent numerical studies suggest that the production rate through this direct ejection is at least two orders of magnitude lower than what is needed to explain the observed number and fraction in the Trapezium Cluster (Yu & Lai, 2024; Portegies Zwart & Hochart, 2023). Another proposed mechanism is through tidal capture during repeated planet-planet scattering in planetary systems (Ochiai et al., 2014; Lazzoni et al., 2023). However, tides generally produce tight binaries rather than wide binaries.

Given the difficulties to form wide binary planets out of the protoplanetary disk (i.e., planet-like formation), star-like formation, in which binary planets form directly within the cluster, may be preferred (see also Portegies Zwart & Hochart, 2023), although the exact formation mechanism remains unclear and requires further investigation.

Once formed, the binary planets in the cluster may evolve as a result of dynamical interactions with passing stars. This is especially the case for such wide binaries as the JWST JuMBOs in the Trapezium Cluster. Due to the extremely high star number density (n1×104pc3similar-tosubscript𝑛1superscript104superscriptpc3n_{\star}\sim 1\times 10^{4}\ \text{pc}^{-3}italic_n start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT ∼ 1 × 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT pc start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT), the dynamical lifetime of a typical JWST JuMBO is comparable to the age of the cluster (see Section 2; see also Portegies Zwart & Hochart 2023). Therefore, it is probable that the wide binary planets seen by JWST were born with closer separations and then got softened by stellar flybys. These binary planets with initially close separations may then be the lower mass counterparts of binaries with L-, T-, and Y-type primaries, for which the binary fraction is 10–30% and the binary separation peaks at 3–10 au (Burgasser, 2007; Joergens, 2008).

Regardless of the true nature of the recently discovered PMOs and JuMBOs by JWST, it is necessary to study the dynamical evolution of the planetary binaries in a dense cluster environment to explore their origins.

2 Analytical Method

The dynamical evolution of binary stars in star clusters has been studied extensively in the literature (e.g., Heggie 1975, Heggie & Hut 2003, and Binney & Tremaine 2008). The problem of the binary planet–star interaction is a simplified version, as the binary mass and binding energy are too small to influence the star’s motion in a meaningful way. Therefore, we aim to build a simple analytical model that focuses on the mMmuch-less-than𝑚subscript𝑀m\ll M_{\star}italic_m ≪ italic_M start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT interaction, where m𝑚mitalic_m is the total mass of the binary and Msubscript𝑀M_{\star}italic_M start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT is the much more massive third body. For simplicity, only equal-mass binaries are considered in this work.

2.1 Effect of a Single Flyby

Refer to caption
Figure 1: Hyperbolic trajectory of the binary planet relative to the passing star. Two planets (blue circles) with a total mass m𝑚mitalic_m orbit their mutual barycenter (black cross) with the relative semimajor axis a𝑎aitalic_a. The periastron distance (closest approach) of the barycenter is denoted as qhsubscript𝑞q_{h}italic_q start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT, which is related to the hyperbolic semimajor axis ahsubscript𝑎a_{h}italic_a start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT and eccentricity ehsubscript𝑒e_{h}italic_e start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT. The geometry of the flyby trajectory is determined by the impact parameter b𝑏bitalic_b, the relative velocity between the binary barycenter and the star at infinity vsubscript𝑣v_{\infty}italic_v start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT, and the stellar mass Msubscript𝑀M_{\star}italic_M start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT.

As illustrated in Figure 1, the hyperbolic trajectory of the binary barycenter relative to the passing star can be described with the following orbital elements:

ahsubscript𝑎\displaystyle a_{h}italic_a start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT =𝒢Mv2,absent𝒢subscript𝑀superscriptsubscript𝑣2\displaystyle=\frac{\mathcal{G}M_{\star}}{v_{\infty}^{2}},= divide start_ARG caligraphic_G italic_M start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG italic_v start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG , (1)
qhsubscript𝑞\displaystyle q_{h}italic_q start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT =ah(eh1)=ah(1+jh21),absentsubscript𝑎subscript𝑒1subscript𝑎1superscriptsubscript𝑗21\displaystyle=a_{h}(e_{h}-1)=a_{h}\left(\sqrt{1+j_{h}^{2}}-1\right),= italic_a start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ( italic_e start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT - 1 ) = italic_a start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ( square-root start_ARG 1 + italic_j start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG - 1 ) ,
jhsubscript𝑗\displaystyle j_{h}italic_j start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT =bah,absent𝑏subscript𝑎\displaystyle=\frac{b}{a_{h}},= divide start_ARG italic_b end_ARG start_ARG italic_a start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT end_ARG ,

where 𝒢𝒢\mathcal{G}caligraphic_G is the gravitational constant. The hyperbolic semimajor axis ahsubscript𝑎a_{h}italic_a start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT and the periastron distance qhsubscript𝑞q_{h}italic_q start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT are determined by the impact parameter b𝑏bitalic_b, the relative velocity between the star and the binary barycenter at infinity vsubscript𝑣v_{\infty}italic_v start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT, and the mass of the passing star Msubscript𝑀M_{\star}italic_M start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT. Assuming M=1Msubscript𝑀1subscript𝑀direct-productM_{\star}=1M_{\odot}italic_M start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT = 1 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, the hyperbolic semimajor axis can be quickly estimated with ah=(30kms1/v)2ausubscript𝑎superscript30kmsuperscripts1subscript𝑣2aua_{h}=(30~{}\text{km}~{}\text{s}^{-1}/v_{\infty})^{2}~{}\text{au}italic_a start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT = ( 30 km s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT / italic_v start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT au. For young massive clusters with a typical velocity dispersion of σ1kms1similar-to𝜎1kmsuperscripts1\sigma\sim 1~{}\text{km}~{}\text{s}^{-1}italic_σ ∼ 1 km s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, ahsubscript𝑎a_{h}italic_a start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT is of the order of similar-to\sim1000 au for solar-mass stars. The ratio between b𝑏bitalic_b and ahsubscript𝑎a_{h}italic_a start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT is defined as jhsubscript𝑗j_{h}italic_j start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT, representing the dimensionless angular momentum on a given semimajor axis (see Tremaine, 2023).

The pericenter velocity vqsubscript𝑣𝑞v_{q}italic_v start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT can be obtained via the conservation of angular momentum

vq=bvqh.subscript𝑣𝑞𝑏subscript𝑣subscript𝑞v_{q}=\frac{bv_{\infty}}{q_{h}}.italic_v start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT = divide start_ARG italic_b italic_v start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT end_ARG start_ARG italic_q start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT end_ARG . (2)

The typical timescale of the encounter is 111There are also other definitions for the encounter timescale, such as b/v𝑏subscript𝑣b/v_{\infty}italic_b / italic_v start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT and qh/vsubscript𝑞subscript𝑣q_{h}/v_{\infty}italic_q start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT / italic_v start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT. However, we notice that these definitions differ significantly from the real interaction timescale, when the trajectory is near-parabolic (eh1similar-tosubscript𝑒1e_{h}\sim 1italic_e start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ∼ 1).

tenc2qhvq,subscript𝑡enc2subscript𝑞subscript𝑣𝑞t_{\text{enc}}\equiv\frac{2q_{h}}{v_{q}},italic_t start_POSTSUBSCRIPT enc end_POSTSUBSCRIPT ≡ divide start_ARG 2 italic_q start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT end_ARG start_ARG italic_v start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT end_ARG , (3)

and a close encounter is defined as tencPless-than-or-similar-tosubscript𝑡enc𝑃t_{\text{enc}}\lesssim Pitalic_t start_POSTSUBSCRIPT enc end_POSTSUBSCRIPT ≲ italic_P, with the orbital period of the binary P𝑃Pitalic_P given by

P=2πa3/(𝒢m).𝑃2𝜋superscript𝑎3𝒢𝑚P=2\pi\sqrt{a^{3}/(\mathcal{G}m)}.italic_P = 2 italic_π square-root start_ARG italic_a start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT / ( caligraphic_G italic_m ) end_ARG . (4)

Here a𝑎aitalic_a is the binary semimajor axis and m𝑚mitalic_m is the binary total mass. During such encounters, the relative location of the two planets does not change over the span when the gravitational pull from the star is most intense. As a result, the change in the specific orbital energy of the binary, E=𝒢m/(2a)𝐸𝒢𝑚2𝑎E=-\mathcal{G}m/(2a)italic_E = - caligraphic_G italic_m / ( 2 italic_a ), can be estimated with an impulse approximation (Farinella & Chauvineau, 1993):

ΔE𝒢2MmP4abvqhflflΨdf.similar-to-or-equalsΔ𝐸superscript𝒢2subscript𝑀𝑚𝑃4𝑎𝑏subscript𝑣subscript𝑞superscriptsubscriptsubscript𝑓𝑙subscript𝑓𝑙Ψdifferential-d𝑓\displaystyle\Delta E\simeq\frac{\mathcal{G}^{2}M_{\star}mP}{4abv_{\infty}q_{h% }}\int_{-f_{l}}^{f_{l}}\Psi\mathrm{d}f.roman_Δ italic_E ≃ divide start_ARG caligraphic_G start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT italic_m italic_P end_ARG start_ARG 4 italic_a italic_b italic_v start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT italic_q start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT end_ARG ∫ start_POSTSUBSCRIPT - italic_f start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_f start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT end_POSTSUPERSCRIPT roman_Ψ roman_d italic_f . (5)

Here, the integral term evaluates the total energy change along the hyperbolic path, which is related to the spatial orientation of the hyperbola relative to the binary orbital plane (ascending node ΩΩ\Omegaroman_Ω and inclination i𝑖iitalic_i) as well as the exact locations of the two components. Monte-Carlo simulations of random flyby angles and initial locations show that the orientation integral has an average of zero and a root mean square of the order of unity (Farinella & Chauvineau, 1993). This implies that random close encounters have equal probabilities of softening and hardening the binary. Therefore, the effect of multiple flybys can be modeled as a random walk in E𝐸Eitalic_E space, with the typical energy change |ΔE|Δ𝐸|\Delta E|| roman_Δ italic_E | (i.e., the variance of the ΔEΔ𝐸\Delta Eroman_Δ italic_E distribution) being

|ΔE|Δ𝐸\displaystyle\left|\Delta E\right|| roman_Δ italic_E | 𝒢2MmP4abvqh.similar-to-or-equalsabsentsuperscript𝒢2subscript𝑀𝑚𝑃4𝑎𝑏subscript𝑣subscript𝑞\displaystyle\simeq\frac{\mathcal{G}^{2}M_{\star}mP}{4abv_{\infty}q_{h}}.≃ divide start_ARG caligraphic_G start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT italic_m italic_P end_ARG start_ARG 4 italic_a italic_b italic_v start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT italic_q start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT end_ARG . (6)

We define critical flybys as stellar encounters with a typical energy change larger than the binary binding energy |ΔE|>|E|Δ𝐸𝐸\left|\Delta E\right|>\left|E\right|| roman_Δ italic_E | > | italic_E |. Such encounters would likely induce significant changes to the semimajor axis a𝑎aitalic_a and orbital eccentricity e𝑒eitalic_e of the binary, or even ionize it. Rearranging the expression, one obtains an inequality for two timescales:

tcr2qhvjh<P,subscript𝑡cr2subscript𝑞subscript𝑣subscript𝑗𝑃\displaystyle t_{\text{cr}}\equiv\frac{2q_{h}}{v_{\infty}}j_{h}<P,italic_t start_POSTSUBSCRIPT cr end_POSTSUBSCRIPT ≡ divide start_ARG 2 italic_q start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT end_ARG start_ARG italic_v start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT end_ARG italic_j start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT < italic_P , (7)

where the left-hand side is defined as the critical flyby timescale tcrsubscript𝑡crt_{\text{cr}}italic_t start_POSTSUBSCRIPT cr end_POSTSUBSCRIPT.

The encounter timescale tencsubscript𝑡enct_{\text{enc}}italic_t start_POSTSUBSCRIPT enc end_POSTSUBSCRIPT (Equation 3) can be rewritten as tenc=(2qh/v)(qh/b)subscript𝑡enc2subscript𝑞subscript𝑣subscript𝑞𝑏t_{\text{enc}}=(2q_{h}/v_{\infty})(q_{h}/b)italic_t start_POSTSUBSCRIPT enc end_POSTSUBSCRIPT = ( 2 italic_q start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT / italic_v start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT ) ( italic_q start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT / italic_b ) using Equation (2). The relation between the two timescales, tencsubscript𝑡enct_{\text{enc}}italic_t start_POSTSUBSCRIPT enc end_POSTSUBSCRIPT and tcrsubscript𝑡crt_{\text{cr}}italic_t start_POSTSUBSCRIPT cr end_POSTSUBSCRIPT, is thus given by

tcr=tenc(jh21+jh21),subscript𝑡crsubscript𝑡encsuperscriptsubscript𝑗21superscriptsubscript𝑗21\displaystyle t_{\text{cr}}=t_{\text{enc}}\left(\frac{j_{h}^{2}}{\sqrt{1+j_{h}% ^{2}}-1}\right),italic_t start_POSTSUBSCRIPT cr end_POSTSUBSCRIPT = italic_t start_POSTSUBSCRIPT enc end_POSTSUBSCRIPT ( divide start_ARG italic_j start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG square-root start_ARG 1 + italic_j start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG - 1 end_ARG ) , (8)

and it is trivial to verify that tcr>2tencsubscript𝑡cr2subscript𝑡enct_{\text{cr}}>2t_{\text{enc}}italic_t start_POSTSUBSCRIPT cr end_POSTSUBSCRIPT > 2 italic_t start_POSTSUBSCRIPT enc end_POSTSUBSCRIPT holds for all jhsubscript𝑗j_{h}italic_j start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT. As a result, tcr<Psubscript𝑡cr𝑃t_{\text{cr}}<Pitalic_t start_POSTSUBSCRIPT cr end_POSTSUBSCRIPT < italic_P is always a stronger constraint than tenc<Psubscript𝑡enc𝑃t_{\text{enc}}<Pitalic_t start_POSTSUBSCRIPT enc end_POSTSUBSCRIPT < italic_P; in other words, all critical flybys are close encounters where the impulse approximation is applicable.

Refer to caption
Figure 2: Critical timescale tcrsubscript𝑡crt_{\text{cr}}italic_t start_POSTSUBSCRIPT cr end_POSTSUBSCRIPT (Equation 7) computed on the b𝑏bitalic_b-vsubscript𝑣v_{\infty}italic_v start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT space assuming a solar-mass star. Dashed lines denote the orbital periods of 10 MJsubscript𝑀𝐽M_{J}italic_M start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT JuMBOs with a wide range of semimajor axes a𝑎aitalic_a, of which critical flybys lie on the left-hand side. The solid line represents jh=1subscript𝑗1j_{h}=1italic_j start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT = 1 (eh=2subscript𝑒2e_{h}=\sqrt{2}italic_e start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT = square-root start_ARG 2 end_ARG), which divides the parameter space into the hyperbolic regime (jh>1subscript𝑗1j_{h}>1italic_j start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT > 1, above the line) and the near-parabolic regime (jh<1subscript𝑗1j_{h}<1italic_j start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT < 1, below the line).

We compute the critical timescale for a series of b𝑏bitalic_b and vsubscript𝑣v_{\infty}italic_v start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT using Equation (7) and plot the results in Figure 2. The critical flyby timescale spans several orders of magnitude, from a few days to more than 100 Myr. tcrsubscript𝑡crt_{\text{cr}}italic_t start_POSTSUBSCRIPT cr end_POSTSUBSCRIPT also scales differently given the geometry of the flyby trajectory: Near-parabolic flybys with eh1similar-tosubscript𝑒1e_{h}\sim 1italic_e start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ∼ 1 (jh0similar-tosubscript𝑗0j_{h}\sim 0italic_j start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ∼ 0) affect the binary more effectively than hyperbolic flybys (jh>1subscript𝑗1j_{h}>1italic_j start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT > 1) do, as indicated by a slope change separated by the solid line jh=1subscript𝑗1j_{h}=1italic_j start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT = 1.

To derive the scaling law, one first expands the pericenter distance

qhsubscript𝑞\displaystyle q_{h}italic_q start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT {b22ah if jh<1,b if jh>1,absentcasescontinued-fractionsuperscript𝑏22subscript𝑎 if subscript𝑗1𝑏 if subscript𝑗1\displaystyle\approx\begin{cases}\cfrac{b^{2}}{2a_{h}}\quad&\text{ if }j_{h}<1% ,\\ b\quad&\text{ if }j_{h}>1,\\ \end{cases}≈ { start_ROW start_CELL continued-fraction start_ARG italic_b start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_a start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT end_ARG end_CELL start_CELL if italic_j start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT < 1 , end_CELL end_ROW start_ROW start_CELL italic_b end_CELL start_CELL if italic_j start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT > 1 , end_CELL end_ROW (9)

and then explicitly writes the fractional orbital energy change for different jhsubscript𝑗j_{h}italic_j start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT

|Δ||ΔE||E|={(𝒢M)2Pb3v3 if jh<1,(𝒢MP/2)b2v1 if jh>1.ΔΔ𝐸𝐸casessuperscript𝒢subscript𝑀2𝑃superscript𝑏3superscriptsubscript𝑣3 if subscript𝑗1𝒢subscript𝑀𝑃2superscript𝑏2superscriptsubscript𝑣1 if subscript𝑗1|\Delta\mathcal{E}|\equiv\frac{|\Delta E|}{|E|}=\begin{cases}\left(\mathcal{G}% M_{\star}\right)^{2}Pb^{-3}v_{\infty}^{-3}\quad&\text{ if }j_{h}<1,\\ \left(\mathcal{G}M_{\star}P/2\right)b^{-2}v_{\infty}^{-1}\quad&\text{ if }j_{h% }>1.\end{cases}| roman_Δ caligraphic_E | ≡ divide start_ARG | roman_Δ italic_E | end_ARG start_ARG | italic_E | end_ARG = { start_ROW start_CELL ( caligraphic_G italic_M start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_P italic_b start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT italic_v start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT end_CELL start_CELL if italic_j start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT < 1 , end_CELL end_ROW start_ROW start_CELL ( caligraphic_G italic_M start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT italic_P / 2 ) italic_b start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT italic_v start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT end_CELL start_CELL if italic_j start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT > 1 . end_CELL end_ROW (10)

Equating |Δ|Δ|\Delta\mathcal{E}|| roman_Δ caligraphic_E | with 1 yields

bcr={[(𝒢M)2P]13v1 if jh<1,[𝒢MP/2]12v12 if jh>1,subscript𝑏crcasessuperscriptdelimited-[]superscript𝒢subscript𝑀2𝑃13superscriptsubscript𝑣1 if subscript𝑗1superscriptdelimited-[]𝒢subscript𝑀𝑃212superscriptsubscript𝑣12 if subscript𝑗1b_{\text{cr}}=\begin{cases}\left[\left(\mathcal{G}M_{\star}\right)^{2}P\right]% ^{\frac{1}{3}}v_{\infty}^{-1}\quad&\text{ if }j_{h}<1,\\ \left[\mathcal{G}M_{\star}P/2\right]^{\frac{1}{2}}v_{\infty}^{-\frac{1}{2}}% \quad&\text{ if }j_{h}>1,\end{cases}italic_b start_POSTSUBSCRIPT cr end_POSTSUBSCRIPT = { start_ROW start_CELL [ ( caligraphic_G italic_M start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_P ] start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 3 end_ARG end_POSTSUPERSCRIPT italic_v start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT end_CELL start_CELL if italic_j start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT < 1 , end_CELL end_ROW start_ROW start_CELL [ caligraphic_G italic_M start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT italic_P / 2 ] start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT italic_v start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT end_CELL start_CELL if italic_j start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT > 1 , end_CELL end_ROW (11)

where bcrsubscript𝑏crb_{\text{cr}}italic_b start_POSTSUBSCRIPT cr end_POSTSUBSCRIPT is the impact parameter for critical flybys. Notice bcrv1proportional-tosubscript𝑏crsuperscriptsubscript𝑣1b_{\text{cr}}\propto v_{\infty}^{-1}italic_b start_POSTSUBSCRIPT cr end_POSTSUBSCRIPT ∝ italic_v start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT in the near-parabolic regime and bcrv12proportional-tosubscript𝑏crsuperscriptsubscript𝑣12b_{\text{cr}}\propto v_{\infty}^{-\frac{1}{2}}italic_b start_POSTSUBSCRIPT cr end_POSTSUBSCRIPT ∝ italic_v start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT in the hyperbolic regime, which corresponds to the slope change in Figure 2.

One can easily estimate the orbital stability of a binary planet against a particular flyby by comparing tcrsubscript𝑡crt_{\text{cr}}italic_t start_POSTSUBSCRIPT cr end_POSTSUBSCRIPT with P𝑃Pitalic_P. In Figure 2, we plot orbital periods of 10 MJsubscript𝑀𝐽M_{J}italic_M start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT JuMBOs with a𝑎aitalic_a spanning from 0.1 au to 1000 au in dashed lines, and the areas to the left of these lines indicate flyby parameters that would likely induce significant changes in binding energy. For JWST JuMBOs with a=25𝑎25a=25italic_a = 25–400 au, critical flybys occur at bcr600less-than-or-similar-tosubscript𝑏cr600b_{\text{cr}}\lesssim 600italic_b start_POSTSUBSCRIPT cr end_POSTSUBSCRIPT ≲ 600–6,000 au for v=1subscript𝑣1v_{\infty}=1italic_v start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT = 1 km s1superscripts1\text{s}^{-1}s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, and bcr300less-than-or-similar-tosubscript𝑏cr300b_{\text{cr}}\lesssim 300italic_b start_POSTSUBSCRIPT cr end_POSTSUBSCRIPT ≲ 300–3,000 au for v=2.5subscript𝑣2.5v_{\infty}=2.5italic_v start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT = 2.5 km s1superscripts1\text{s}^{-1}s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT. It is worth pointing out that such close flybys may be happening in some of the observed systems (see JuMBO33 and 34 in Figure 3 of Pearson & McCaughrean 2023)

2.2 Dynamical Lifetime Inside a Cluster

Until now, we have only looked at the average effects caused by a single passing star. However, in a cluster environment where thousands of stars are present, it is essential to consider the ongoing interactions from multiple stars, each with varying trajectories and characteristics.

The frequency of encounters with the parameters b𝑏bitalic_b and vsubscript𝑣v_{\infty}italic_v start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT for a period of time t𝑡titalic_t is given by

N=nt(πb2)v,𝑁subscript𝑛𝑡𝜋superscript𝑏2subscript𝑣N=n_{\star}t(\pi b^{2})v_{\infty},italic_N = italic_n start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT italic_t ( italic_π italic_b start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) italic_v start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT , (12)

where nsubscript𝑛n_{\star}italic_n start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT is the number density of the star cluster, and πb2𝜋superscript𝑏2\pi b^{2}italic_π italic_b start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT represents the cross section of the encounter. For critical flybys with b<bcr𝑏subscript𝑏crb<b_{\text{cr}}italic_b < italic_b start_POSTSUBSCRIPT cr end_POSTSUBSCRIPT, the induced fractional energy change ΔΔ\Delta\mathcal{E}roman_Δ caligraphic_E is of the order of unity. In contrast, the energy drift of non-critical distant flybys can build up, whose cumulative change after N𝑁Nitalic_N flybys follows the random walk equation |Δ|N=N|Δ|subscriptΔ𝑁𝑁Δ|\Delta\mathcal{E}|_{N}=\sqrt{N}|\Delta\mathcal{E}|| roman_Δ caligraphic_E | start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT = square-root start_ARG italic_N end_ARG | roman_Δ caligraphic_E |. Apparently, |Δ|Nb2proportional-tosubscriptΔ𝑁superscript𝑏2|\Delta\mathcal{E}|_{N}\propto b^{-2}| roman_Δ caligraphic_E | start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT ∝ italic_b start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT in the near-parabolic case and b1proportional-toabsentsuperscript𝑏1\propto b^{-1}∝ italic_b start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT in the hyperbolic case. In other words, in both cases, the total energy change is dominated by the single closest stellar encounter, rather than the cumulative effects of more frequent distant encounters. As a result, the ionization of a planetary-mass binary mostly results from the closest encounter, and the dynamical lifetime can be defined as the expected time interval between two critical flybys (Equation 13).

T𝑇\displaystyle Titalic_T 1n(πbcr2)vabsent1subscript𝑛𝜋superscriptsubscript𝑏cr2subscript𝑣\displaystyle\equiv\frac{1}{n_{\star}(\pi b_{\text{cr}}^{2})v_{\infty}}≡ divide start_ARG 1 end_ARG start_ARG italic_n start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT ( italic_π italic_b start_POSTSUBSCRIPT cr end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) italic_v start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT end_ARG (13)
{0.9Myr(n104pc3)1(MM)43(m10MJ)13(a100au)1(v1km s1) if jh<1,1.4Myr(n104pc3)1(MM)1(m10MJ)12(a100au)32 if jh>1.similar-to-or-equalsabsentcases0.9Myrsuperscriptcontinued-fractionsubscript𝑛superscript104superscriptpc31superscriptcontinued-fractionsubscript𝑀subscript𝑀direct-product43superscriptcontinued-fraction𝑚10subscript𝑀𝐽13superscriptcontinued-fraction𝑎100au1continued-fractionsubscript𝑣1superscriptkm s1 if subscript𝑗11.4Myrsuperscriptcontinued-fractionsubscript𝑛superscript104superscriptpc31superscriptcontinued-fractionsubscript𝑀subscript𝑀direct-product1superscriptcontinued-fraction𝑚10subscript𝑀𝐽12superscriptcontinued-fraction𝑎100au32 if subscript𝑗1\displaystyle\simeq\begin{cases}0.9~{}\text{Myr}~{}\left(\cfrac{n_{\star}}{10^% {4}~{}\text{pc}^{-3}}\right)^{-1}\left(\cfrac{M_{\star}}{M_{\odot}}\right)^{-% \frac{4}{3}}\left(\cfrac{m}{10~{}M_{J}}\right)^{\frac{1}{3}}\left(\cfrac{a}{10% 0~{}\text{au}}\right)^{-1}\left(\cfrac{v_{\infty}}{1~{}\text{km s}^{-1}}\right% )\quad&\text{ if }j_{h}<1,\\ 1.4~{}\text{Myr}~{}\left(\cfrac{n_{\star}}{10^{4}~{}\text{pc}^{-3}}\right)^{-1% }\left(\cfrac{M_{\star}}{M_{\odot}}\right)^{-1}\left(\cfrac{m}{10~{}M_{J}}% \right)^{\frac{1}{2}}\left(\cfrac{a}{100~{}\text{au}}\right)^{-\frac{3}{2}}% \quad&\text{ if }j_{h}>1.\end{cases}≃ { start_ROW start_CELL 0.9 Myr ( continued-fraction start_ARG italic_n start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT pc start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT end_ARG ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( continued-fraction start_ARG italic_M start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT - divide start_ARG 4 end_ARG start_ARG 3 end_ARG end_POSTSUPERSCRIPT ( continued-fraction start_ARG italic_m end_ARG start_ARG 10 italic_M start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 3 end_ARG end_POSTSUPERSCRIPT ( continued-fraction start_ARG italic_a end_ARG start_ARG 100 au end_ARG ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( continued-fraction start_ARG italic_v start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT end_ARG start_ARG 1 km s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT end_ARG ) end_CELL start_CELL if italic_j start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT < 1 , end_CELL end_ROW start_ROW start_CELL 1.4 Myr ( continued-fraction start_ARG italic_n start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT pc start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT end_ARG ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( continued-fraction start_ARG italic_M start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( continued-fraction start_ARG italic_m end_ARG start_ARG 10 italic_M start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT ( continued-fraction start_ARG italic_a end_ARG start_ARG 100 au end_ARG ) start_POSTSUPERSCRIPT - divide start_ARG 3 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT end_CELL start_CELL if italic_j start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT > 1 . end_CELL end_ROW

Notice that the discontinuity between the two equations at jh=1subscript𝑗1j_{h}=1italic_j start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT = 1 arises from different approximations of qhsubscript𝑞q_{h}italic_q start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT in Equation (9). Adopting the typical number density of 1×104pc31superscript104superscriptpc31\times 10^{4}~{}\text{pc}^{-3}1 × 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT pc start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT and the velocity dispersion of 1km s11superscriptkm s11~{}\text{km s}^{-1}1 km s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT for young star clusters, Equation (13) suggests that the dynamical lifetime of a JuMBO with a=100𝑎100a=100italic_a = 100 au is only of the order of similar-to\sim1 Myr. It should be noted that the structure of star clusters dynamically evolves, so the JuMBO stability may vary depending on the cluster density over time.

We also calculate the dynamical lifetime of potential JuMBOs within other types of cluster, assuming a constant star density and relative velocity. For open clusters such as Hyades (n=1pc3subscript𝑛1superscriptpc3n_{\star}=1~{}\text{pc}^{-3}italic_n start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT = 1 pc start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT and v=0.25km s1subscript𝑣0.25superscriptkm s1v_{\infty}=0.25~{}\text{km s}^{-1}italic_v start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT = 0.25 km s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, Perryman et al. 1998), binary planets similar to the JWST JuMBOs are stable for several billion years, significantly longer than the similar-to\sim700 Myr age of the Hyades; Conversely, the same type of binary planets would have long been destabilized if they formed in the globular cluster 47 Tucanae (n=1×105pc3subscript𝑛1superscript105superscriptpc3n_{\star}=1\times 10^{5}~{}\text{pc}^{-3}italic_n start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT = 1 × 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT pc start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT and v=10km s1subscript𝑣10superscriptkm s1v_{\infty}=10~{}\text{km s}^{-1}italic_v start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT = 10 km s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, Meylan & Mayor 1986), given their short dynamical lifetime (less-than-or-similar-to\lesssim1 Myr) compared to the cluster age of 11 Gyr.

2.3 Numerical Validation

We carry out numerical simulations in which binary planets encounter a solar-mass star from random directions. For each simulation, one thousand equal-mass binaries with a binary mass of 10MJ10subscript𝑀𝐽10M_{J}10 italic_M start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT are initialized on a circular orbit with a0=15subscript𝑎015a_{0}=15italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 15 au. Their barycenter is then set on a hyperbolic trajectory relative to the passing star with v=1kms1subscript𝑣1kmsuperscripts1v_{\infty}=1~{}\text{km}~{}\text{s}^{-1}italic_v start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT = 1 km s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT. Both the initial binary mean anomalies and the longitude of the ascending nodes are distributed uniformly in (0,2π)02𝜋(0,2\pi)( 0 , 2 italic_π ), while the relative inclinations between the hyperbolic trajectory and the binary orbital plane are distributed uniformly in cos(i)𝑖\cos(i)roman_cos ( italic_i ), corresponding to the isotropic angles of entry. We then obtain the corresponding critical impact parameter bcr=460subscript𝑏cr460b_{\text{cr}}=460italic_b start_POSTSUBSCRIPT cr end_POSTSUBSCRIPT = 460 au using Equation (11), and set b=bcr/1.5,bcr,𝑏subscript𝑏cr1.5subscript𝑏crb=b_{\text{cr}}/1.5,b_{\text{cr}},italic_b = italic_b start_POSTSUBSCRIPT cr end_POSTSUBSCRIPT / 1.5 , italic_b start_POSTSUBSCRIPT cr end_POSTSUBSCRIPT , and 1.5bcr1.5subscript𝑏cr1.5b_{\text{cr}}1.5 italic_b start_POSTSUBSCRIPT cr end_POSTSUBSCRIPT, respectively, for the three experiments. All three cases have jh<1subscript𝑗1j_{h}<1italic_j start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT < 1 and thus belong to the near-parabolic regime. Simulations are performed using the ias15 integrator of rebound (Rein & Spiegel, 2014).

Refer to caption
Figure 3: Upper: Orbital distributions and histograms of 1,000 equal-mass binary planets after encountering a solar-mass star with v=1kms1subscript𝑣1kmsuperscripts1v_{\infty}=1~{}\text{km}~{}\text{s}^{-1}italic_v start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT = 1 km s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT. Binaries are initialized with a0=15subscript𝑎015a_{0}=15italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 15 au, e0=0subscript𝑒00e_{0}=0italic_e start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0, and m=10MJ𝑚10subscript𝑀𝐽m=10~{}M_{J}italic_m = 10 italic_M start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT, while the chosen impact parameters correspond to a deep critical flyby (bcr/1.5subscript𝑏cr1.5b_{\text{cr}}/1.5italic_b start_POSTSUBSCRIPT cr end_POSTSUBSCRIPT / 1.5, left panel), a critical flyby (bcrsubscript𝑏crb_{\text{cr}}italic_b start_POSTSUBSCRIPT cr end_POSTSUBSCRIPT, middle panel), and a distant flyby (1.5bcr1.5subscript𝑏cr1.5b_{\text{cr}}1.5 italic_b start_POSTSUBSCRIPT cr end_POSTSUBSCRIPT, right panel), respectively. The x𝑥xitalic_x axis indicates the fractional energy change ΔΔ\Delta\mathcal{E}roman_Δ caligraphic_E. The left y𝑦yitalic_y axis represents the binary eccentricity post flyby, whereas the right y𝑦yitalic_y axis denotes the number of binaries in each bin. The orange dashed line marks the energy change of 0, with a negative ΔΔ\Delta\mathcal{E}roman_Δ caligraphic_E corresponding to a decrease in a𝑎aitalic_a, and vice versa (red arrows). The green shaded region denotes the observable JuMBOs with a>25𝑎25a>25italic_a > 25 au. Lower: Histograms of semimajor axes of captured planets (acapsubscript𝑎capa_{\text{cap}}italic_a start_POSTSUBSCRIPT cap end_POSTSUBSCRIPT) from ionized JuMBO pairs.

The post-flyby orbital distributions of the three simulations are shown in Figure 3, with ΔΔ\Delta\mathcal{E}roman_Δ caligraphic_E denoting the fractional energy change and e𝑒eitalic_e denoting the eccentricity. We decide to plot ΔΔ\Delta\mathcal{E}roman_Δ caligraphic_E instead of ΔaΔ𝑎\Delta aroman_Δ italic_a because it is a dimensionless scale-free quantity. In other words, even if we ran the same simulations (with corresponding bcrsubscript𝑏crb_{\text{cr}}italic_b start_POSTSUBSCRIPT cr end_POSTSUBSCRIPT) for binaries with different m𝑚mitalic_m and a0subscript𝑎0a_{0}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, their orbital distributions in the ΔΔ\Delta\mathcal{E}roman_Δ caligraphic_E space would still look similar to those in Figure 3.

We define the following branching fractions:

  1. 1.

    f<25subscript𝑓absent25f_{<\text{25}}italic_f start_POSTSUBSCRIPT < 25 end_POSTSUBSCRIPT — the number of binaries with a<25𝑎25a<25italic_a < 25 au divided by the total number of primordial binaries,

  2. 2.

    f>25subscript𝑓absent25f_{>\text{25}}italic_f start_POSTSUBSCRIPT > 25 end_POSTSUBSCRIPT — the number of binaries with a>25𝑎25a>25italic_a > 25 au divided by the total number, and

  3. 3.

    fionsubscript𝑓ionf_{\text{ion}}italic_f start_POSTSUBSCRIPT ion end_POSTSUBSCRIPT — the number of ionized binaries (signified by a positive binary energy post flyby) divided by the total number (i.e., ionization fraction).

With the above definitions, we have f<25+f>25+fion=1subscript𝑓absent25subscript𝑓absent25subscript𝑓ion1f_{<\text{25}}+f_{>\text{25}}+f_{\text{ion}}=1italic_f start_POSTSUBSCRIPT < 25 end_POSTSUBSCRIPT + italic_f start_POSTSUBSCRIPT > 25 end_POSTSUBSCRIPT + italic_f start_POSTSUBSCRIPT ion end_POSTSUBSCRIPT = 1. The values of these three fractions are shown in Figure 3.

The left and middle panels of Figure 3 demonstrate the aftermath of critical flybys with different impact parameters. Although both flybys show roughly equal chances of softening and hardening the orbit (that is, a near symmetric distribution centered at 00), the middle panel shows a strong peak at Δ=0Δ0\Delta\mathcal{E}=0roman_Δ caligraphic_E = 0, implying that a significant portion of binaries are not strongly affected by the flyby. In comparison, 64% of the binaries are ionized by deeper critical flybys.

When the equal-mass binaries are ionized by the passing star, a common outcome is the capture of one planet by the passing star and the ejection of the other planet222Additional simulations with unequal-mass binaries were also carried out, and the result suggests a lower capture probability.. The captured planets are on highly eccentric orbits around the star, with acapsubscript𝑎capa_{\text{cap}}italic_a start_POSTSUBSCRIPT cap end_POSTSUBSCRIPT peaked at 300–400 au and the minimal eccentricity of emin=0.5subscript𝑒min0.5e_{\text{min}}=0.5italic_e start_POSTSUBSCRIPT min end_POSTSUBSCRIPT = 0.5–0.6 (lower panels in Figure 3). These distant and elliptic planetary orbits are mostly not stable inside a dense cluster environment, and one can expect that subsequent stellar perturbations would easily disrupt them and produce FFPs (Spurzem et al., 2009). It is worth noting that three-body interactions occurred frequently in the early Solar System, both among planetesimals (Funato et al., 2004) and between planetesimals and a planet. For example, Triton – the largest moon of Neptune – is thought to be captured during a close encounter between Neptune and a transneptunian binary (Agnor & Hamilton, 2006).

There is an apparent dichotomy between critical flybys and noncritical flybys. As illustrated in the right panel of Figure 3, the noncritical (b=1.5bcr𝑏1.5subscript𝑏crb=1.5b_{\text{cr}}italic_b = 1.5 italic_b start_POSTSUBSCRIPT cr end_POSTSUBSCRIPT) flybys induce little variation in both the binary energy and eccentricity. In a cluster environment, such flybys would be roughly twice as likely as critical (bcrsubscript𝑏crb_{\text{cr}}italic_b start_POSTSUBSCRIPT cr end_POSTSUBSCRIPT) flybys (Equation 12). However, the cumulative effect of two 1.5bcr1.5subscript𝑏cr1.5b_{\text{cr}}1.5 italic_b start_POSTSUBSCRIPT cr end_POSTSUBSCRIPT flybys is by no means comparable to a bcrsubscript𝑏crb_{\text{cr}}italic_b start_POSTSUBSCRIPT cr end_POSTSUBSCRIPT flyby. This again validates our previous analysis that the orbital change of a binary planet is dominated by the single closest passing star.

3 Application to JWST JuMBOs

We have demonstrated that starting from a single population, gravitational interactions with passing stars are able to produce a variety of binary systems from a few to several hundred au. The same dynamical process also produces planetary binaries on tighter orbits (close binaries), FFPs, and temporarily captured planets, which are expected to be liberated as FFPs shortly after due to their extremely large semimajor axes.

Limited by the observational capability of the telescope, however, only wide binaries with separations above a certain threshold (25252525 au, for JuMBOs in the Trapezium) can be resolved by JWST, and it is unclear whether the ‘non-binary’ PMOs are isolated FFPs, or unresolved close JuMBOs, or more likely, a combination of both. The applicable constraint from Pearson & McCaughrean (2023) is the similar-to\sim9% JuMBO-PMO fraction (the number of a>25𝑎25a>25italic_a > 25 au JuMBOs divided by the total number of all planetary-mass objects), i.e., the wide binary fraction fwidesubscript𝑓widef_{\text{wide}}italic_f start_POSTSUBSCRIPT wide end_POSTSUBSCRIPT, which is also given by

fwidesubscript𝑓wide\displaystyle f_{\text{wide}}italic_f start_POSTSUBSCRIPT wide end_POSTSUBSCRIPT =f>252×fion+f<25+f>25+RS:Babsentsubscript𝑓absent252subscript𝑓ionsubscript𝑓absent25subscript𝑓absent25subscript𝑅:𝑆𝐵\displaystyle=\frac{f_{>25}}{2\times f_{\text{ion}}+f_{<25}+f_{>25}+R_{S:B}}= divide start_ARG italic_f start_POSTSUBSCRIPT > 25 end_POSTSUBSCRIPT end_ARG start_ARG 2 × italic_f start_POSTSUBSCRIPT ion end_POSTSUBSCRIPT + italic_f start_POSTSUBSCRIPT < 25 end_POSTSUBSCRIPT + italic_f start_POSTSUBSCRIPT > 25 end_POSTSUBSCRIPT + italic_R start_POSTSUBSCRIPT italic_S : italic_B end_POSTSUBSCRIPT end_ARG (14)
=f>251+fion+RS:B,absentsubscript𝑓absent251subscript𝑓ionsubscript𝑅:𝑆𝐵\displaystyle=\frac{f_{>25}}{1+f_{\text{ion}}+R_{S:B}},= divide start_ARG italic_f start_POSTSUBSCRIPT > 25 end_POSTSUBSCRIPT end_ARG start_ARG 1 + italic_f start_POSTSUBSCRIPT ion end_POSTSUBSCRIPT + italic_R start_POSTSUBSCRIPT italic_S : italic_B end_POSTSUBSCRIPT end_ARG ,

where RS:B=Nsingle/Nbinarysubscript𝑅:𝑆𝐵subscript𝑁singlesubscript𝑁binaryR_{S:B}=N_{\text{single}}/N_{\text{binary}}italic_R start_POSTSUBSCRIPT italic_S : italic_B end_POSTSUBSCRIPT = italic_N start_POSTSUBSCRIPT single end_POSTSUBSCRIPT / italic_N start_POSTSUBSCRIPT binary end_POSTSUBSCRIPT is the primordial ratio between single planets and binary planets governed by the formation process. Since the formation mechanism is still unknown, we simply assume that it only produces binary objects, that is, RS:B=0subscript𝑅:𝑆𝐵0R_{S:B}=0italic_R start_POSTSUBSCRIPT italic_S : italic_B end_POSTSUBSCRIPT = 0. As a result, Equation 14 gives the upper bound of fwidesubscript𝑓widef_{\text{wide}}italic_f start_POSTSUBSCRIPT wide end_POSTSUBSCRIPT.

The ionization fractions for the two critical flybys presented in Figure 3 are 64% and 27%, respectively. Deeper critical flybys are also more efficient in making a>25𝑎25a>25italic_a > 25 au binaries (6.3% compared to 2.0%), because of stronger energy kicks. The resultant fwidesubscript𝑓widef_{\text{wide}}italic_f start_POSTSUBSCRIPT wide end_POSTSUBSCRIPT therefore ranges from 2% to 4%. It is worth pointing out that fwidesubscript𝑓widef_{\text{wide}}italic_f start_POSTSUBSCRIPT wide end_POSTSUBSCRIPT is very sensitive to the initial a0subscript𝑎0a_{0}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and the definition of wide JuMBOs. Therefore, we next explore numerically what initial a0subscript𝑎0a_{0}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT best reproduces the similar-to\sim9% wide binary fraction, assuming a “primordial JuMBO population” formed at the beginning of the star cluster.

Additional three-body simulations similar to those in Figure 3 were performed. Instead of fixing b𝑏bitalic_b, we initialize hyperbolic trajectories based on a realistic distribution (dN/dbbproportional-tod𝑁d𝑏𝑏\mathrm{d}N/\mathrm{d}b\propto broman_d italic_N / roman_d italic_b ∝ italic_b), with the maximum b𝑏bitalic_b being what gives N=1𝑁1N=1italic_N = 1 in Equation (12). To explore different cluster environments, we define a single parameter χ=0tndt𝜒superscriptsubscript0𝑡subscript𝑛differential-d𝑡\chi=\int_{0}^{t}n_{\star}\mathrm{d}titalic_χ = ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT italic_n start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT roman_d italic_t as the stellar number density-weighted residence time (Batygin et al., 2020), which indicates the time-accumulated effect taking into account the evolution of the cluster. In the lowest order, χnt𝜒subscript𝑛𝑡\chi\approx n_{\star}titalic_χ ≈ italic_n start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT italic_t assuming a fixed star density after cluster formation.

Refer to caption
Figure 4: Ionization fraction fionsubscript𝑓ionf_{\text{ion}}italic_f start_POSTSUBSCRIPT ion end_POSTSUBSCRIPT (+++) and wide binary fraction fwidesubscript𝑓widef_{\text{wide}}italic_f start_POSTSUBSCRIPT wide end_POSTSUBSCRIPT (×\times×) from stellar flyby simulations as a function of the binary period P𝑃Pitalic_P. Each symbol represents the result of 5,000 JuMBOs with the same mass (from right to left, m=1,2,5𝑚125m=1,2,5italic_m = 1 , 2 , 5, and 10 MJsubscript𝑀𝐽M_{J}italic_M start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT, respectively) and a0subscript𝑎0a_{0}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (grouped by the same color) interacting with a solar-mass star. The impact parameter b𝑏bitalic_b is generated from a realistic distribution assuming a density-weighted residence time of χ=2×104𝜒2superscript104\chi=2\times 10^{4}italic_χ = 2 × 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT Myr pc-3 and a fixed relative velocity vsubscript𝑣v_{\infty}italic_v start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT of 1kms11kmsuperscripts11~{}\text{km}~{}\text{s}^{-1}1 km s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT. Ncrsubscript𝑁crN_{\text{cr}}italic_N start_POSTSUBSCRIPT cr end_POSTSUBSCRIPT on the top of the x𝑥xitalic_x axis marks orbital periods where the frequency of critical flybys is >>>0.10.10.10.1 and >>>1111, respectively. The range of fwidesubscript𝑓widef_{\text{wide}}italic_f start_POSTSUBSCRIPT wide end_POSTSUBSCRIPT potentially consistent with the JWST discovery (1%–10%) is highlighted in yellow.

In Figure 4, pluses and crosses represent fionsubscript𝑓ionf_{\text{ion}}italic_f start_POSTSUBSCRIPT ion end_POSTSUBSCRIPT and fwidesubscript𝑓widef_{\text{wide}}italic_f start_POSTSUBSCRIPT wide end_POSTSUBSCRIPT from a simulation set in which 5,000 binary planets encounter random passing stars. The x𝑥xitalic_x axis denotes the binary period P𝑃Pitalic_P, a proxy for its dynamical stability (see Equation 7). The chosen χ=2×104𝜒2superscript104\chi=2\times 10^{4}italic_χ = 2 × 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT Myr pc-3 environment correspond to the lower bound of the Trapezium Cluster core density (Heggie & Aarseth, 1992; Hillenbrand & Hartmann, 1998). As shown in both panels, the ionization fraction fionsubscript𝑓ionf_{\text{ion}}italic_f start_POSTSUBSCRIPT ion end_POSTSUBSCRIPT grows monotonically as the binary period, in agreement with analytical theory.

For primordially tight JuMBOs (a0<25subscript𝑎025a_{0}<25italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT < 25 au), orbital expansion is the main cause for generating wide binaries ranging from 25 to 400 au. The post-evolution fwidesubscript𝑓widef_{\text{wide}}italic_f start_POSTSUBSCRIPT wide end_POSTSUBSCRIPT lies between 1% to 10% (yellow highlight) for a0=10,15subscript𝑎01015a_{0}=10,15italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 10 , 15, and 20 au, consistent with the similar-to\sim9% observed by JWST in terms of order of magnitude. fwidesubscript𝑓widef_{\text{wide}}italic_f start_POSTSUBSCRIPT wide end_POSTSUBSCRIPT is merely similar-to\sim0.1% for a0=5subscript𝑎05a_{0}=5italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 5 au JuMBOs, due to stronger orbital stability (signified by their low ionization fractions) and fewer chances of getting wide orbits. These two effects can be quantitatively estimated in the following manner: 1) To increase a binary a𝑎aitalic_a from 5 au to >>>25 au, a fractional energy change of Δ=(0.8,1)Δ0.81\Delta\mathcal{E}=(0.8,1)roman_Δ caligraphic_E = ( 0.8 , 1 ) is required. Compared to Δ=(0.4,1)Δ0.41\Delta\mathcal{E}=(0.4,1)roman_Δ caligraphic_E = ( 0.4 , 1 ) needed for raising 15 au to >>>25 au, it is an event at least a factor of three (0.6/0.2) less likely to occur with random energy kicks. 2) Critical flybys are roughly 5×5\times5 × more frequent for a0=15subscript𝑎015a_{0}=15italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 15 au than for 5555 au (obtained by combining Equations 11 and 12). Therefore, we expect a factor of similar-to\sim15 difference in fwidesubscript𝑓widef_{\text{wide}}italic_f start_POSTSUBSCRIPT wide end_POSTSUBSCRIPT for a0=15subscript𝑎015a_{0}=15italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 15 au and 5555 au, in excellent agreement with the left panel of Figure 4. There is also a factor of similar-to\sim2 difference in fwidesubscript𝑓widef_{\text{wide}}italic_f start_POSTSUBSCRIPT wide end_POSTSUBSCRIPT for JuMBOs with different masses: more massive JuMBOs tend to have a smaller wide binary fraction, in line with JWST observation (Figure 4 in Pearson & McCaughrean, 2023). In contrast, Equation (13) shows that if JuMBOs primordially formed as wide binaries (a0>25subscript𝑎025a_{0}>25italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT > 25 au), the more massive JuMBOs (which are more stable against stellar flybys) would tend to have higher fwidesubscript𝑓widef_{\text{wide}}italic_f start_POSTSUBSCRIPT wide end_POSTSUBSCRIPT, contradicting the JWST observation.

Therefore, we conclude that if a primordial JuMBO population formed within the Trapezium Cluster, the scenario in which they were initially on tight orbits (a0=10subscript𝑎010a_{0}=10italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 10–20 au) reproduces the observed ratio. This scenario requires a high density-weighted residence time of χ104greater-than-or-equivalent-to𝜒superscript104\chi\gtrsim 10^{4}italic_χ ≳ 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT Myr pc-3, suggesting primordial JuMBOs must have spent most of the time near the core of the cluster until critical flybys significantly changed their orbital properties. This postulated population may also connect to tighter JuMBOs (a<5𝑎5a<5italic_a < 5 au, see Beichman et al. 2013 and Best et al. 2017), and our results indicate that the current PMO population observed by JWST is partially made up of unresolved binaries. Follow-up observations that characterize the properties of these PMOs are the key to unveiling their origins.

4 Discussion

In this work, we have built analytical tools to study the orbital evolution and dynamical stability of binary planets within star clusters. By conducting three-body simulations of random encounters between binary planets and stars, we have shown that critical stellar flybys are efficient at producing binary planets with a wide range of a𝑎aitalic_a and e𝑒eitalic_e from a dominant population.

Applying our results to recently discovered JWST JuMBOs (McCaughrean & Pearson, 2023), we show that their typical dynamical lifetime is comparable to the age of the Trapezium cluster, suggesting that these wide binaries may have encountered stellar flybys and thus dynamically evolved. We propose that the JWST JuMBOs were formed as tighter binaries near/within the core of the Trapezium cluster. To best reproduce the observed similar-to\sim9% wide binary fraction, an initial semimajor axis of a0=10subscript𝑎010a_{0}=10italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 10–20 au and a density-weighted residence time of χ104greater-than-or-equivalent-to𝜒superscript104\chi\gtrsim 10^{4}italic_χ ≳ 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT Myr pc-3 are favored.

The discovered JWST JuMBOs are currently within a radius of similar-to\sim0.6 pc, larger than the core of the Trapezium cluster, which has a radius of similar-to\sim0.1–0.2 pc and a central density of similar-to\sim2–5×104absentsuperscript104\times 10^{4}× 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT pc-3 (McCaughrean & Stauffer, 1994; Hillenbrand & Hartmann, 1998).

Given that the dynamical timescale of the ONC (as well as the inner Trapezium Cluster) is comparable to its age, it is likely that the dynamical equilibrium of the cluster was either recently reached or not yet established (Portegies Zwart et al., 2010). Furthermore, three-body simulations of young massive clusters suggest that the stars were more clustered at an early stage, leading to a higher stellar density (see Figure 1 of Fujii, 2015). Taking into account the dynamical evolution of the cluster itself, it is thus reasonable to postulate that the discovered JWST JuMBOs have experienced χ104greater-than-or-equivalent-to𝜒superscript104\chi\gtrsim 10^{4}italic_χ ≳ 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT Myr pc-3 in the past similar-to\sim1 Myr.

The current JWST JuMBO distribution beyond the cluster core could also be produced by binary–star encounters: JuMBOs that had critical stellar encounters in the past are also those that had experienced a significant barycentric velocity kick. Specifically, the 90 deflection impact parameter b90subscript𝑏90b_{\text{90}}italic_b start_POSTSUBSCRIPT 90 end_POSTSUBSCRIPT has exactly the same definition as ahsubscript𝑎a_{h}italic_a start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT (Binney & Tremaine, 2008), suggesting that any stellar encounters in the near-parabolic regime (b<ah𝑏subscript𝑎b<a_{h}italic_b < italic_a start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT) would also deflect the relative velocity vector by >>>90. This would induce a barycentric velocity change comparable to its original speed. Therefore, even in a scenario where JuMBOs are hypothesized to have formed closer to the core, it is not so surprising to discover escaped members within 1 pc around the core at 1 Myr, given the typical barycentric velocity of v=1km s11pc Myr1𝑣1superscriptkm s11superscriptpc Myr1v=1~{}\text{km s}^{-1}\approx 1~{}\text{pc Myr}^{-1}italic_v = 1 km s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ≈ 1 pc Myr start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT. Future optical surveys on the outskirts of the Trapezium Cluster should help test this validity of this scenario.

During the encounters between binary planets and the flyby stars, one of the planets in the binary may get captured by the star into a wide-separation and eccentric orbit. Our simulations show that the production rate of such captured planets is not negligible, especially in the dense environment. This can potentially provide an alternative explanation for the existence of Jupiter-mass planets on extremely wide and eccentric orbits that have been found by direct imaging surveys (e.g., Kraus et al., 2014; Bryan et al., 2016). It may also be connected to the temporarily present rogue planet (Gladman & Chan, 2006; Huang et al., 2022), or the hypothesized “Planet Nine” (Batygin et al., 2019), in the solar system.

Acknowledgments

We thank Subo Dong and Zenghua Zhang for useful discussions. Y.H. acknowledges funding support from Tsinghua University and NAOJ. This work is supported by the National Science Foundation of China (Grant No. 12173021 and 12133005) and the CASSACA grant CCJRF2105. E. K. is supported by JSPS KAKENHI Grant Number 18H05438. We also acknowledge the Tsinghua Astrophysics High-Performance Computing platform for providing computational and data storage resources.

Software: Rebound (Rein & Liu, 2012), Numpy (Harris et al., 2020), Matplotlib (Hunter, 2007)

References

  • Agnor & Hamilton (2006) Agnor, C. B., & Hamilton, D. P. 2006, Nature, 441, 192, doi: 10.1038/nature04792
  • Batygin et al. (2020) Batygin, K., Adams, F. C., Batygin, Y. K., & Petigura, E. A. 2020, AJ, 159, 101, doi: 10.3847/1538-3881/ab665d
  • Batygin et al. (2019) Batygin, K., Adams, F. C., Brown, M. E., & Becker, J. C. 2019, Phys. Rep., 1 , doi: 10.1016/j.physrep.2019.01.009
  • Beichman et al. (2013) Beichman, C., Gelino, C. R., Kirkpatrick, J. D., et al. 2013, ApJ, 764, 101, doi: 10.1088/0004-637x/764/1/101
  • Best et al. (2017) Best, W. M. J., Liu, M. C., Dupuy, T. J., & Magnier, E. A. 2017, ApJ, 843, L4, doi: 10.3847/2041-8213/aa76df
  • Binney & Tremaine (2008) Binney, J., & Tremaine, S. 2008, Galactic Dynamics: Second Edition (Princeton, NJ USA: Princeton University Press)
  • Bryan et al. (2016) Bryan, M. L., Bowler, B. P., Knutson, H. A., et al. 2016, ApJ, 827, 100, doi: 10.3847/0004-637x/827/2/100
  • Burgasser (2007) Burgasser, A. J. 2007, ApJ, 659, 655, doi: 10.1086/511027
  • Farinella & Chauvineau (1993) Farinella, P., & Chauvineau, B. 1993, A&A, 279, 251
  • Fujii (2015) Fujii, M. S. 2015, Publications of the Astronomical Society of Japan, 67, 59, doi: 10.1093/pasj/psu137
  • Funato et al. (2004) Funato, Y., Makino, J., Hut, P., Kokubo, E., & Kinoshita, D. 2004, Nature, 427, 518, doi: 10.1038/nature02323
  • Gladman & Chan (2006) Gladman, B., & Chan, C. 2006, ApJ, 643, L135, doi: 10.1086/505214
  • Harris et al. (2020) Harris, C. R., Millman, K. J., Walt, S. J. v. d., et al. 2020, Nature, 585, 357, doi: 10.1038/s41586-020-2649-2
  • Heggie & Hut (2003) Heggie, D., & Hut, P. 2003, The Gravitational Million-Body Problem: A Multidisciplinary Approach to Star Cluster Dynamics (Cambridge University Press), doi: 10.1017/cbo9781139164535.023
  • Heggie (1975) Heggie, D. C. 1975, MNRAS, 173, 729, doi: 10.1093/mnras/173.3.729
  • Heggie & Aarseth (1992) Heggie, D. C., & Aarseth, S. J. 1992, MNRAS, 257, 513, doi: 10.1093/mnras/257.3.513
  • Hillenbrand & Hartmann (1998) Hillenbrand, L. A., & Hartmann, L. W. 1998, ApJ, 492, 540, doi: 10.1086/305076
  • Huang et al. (2022) Huang, Y., Gladman, B., Beaudoin, M., & Zhang, K. 2022, ApJ, 938, L23, doi: 10.3847/2041-8213/ac9480
  • Hunter (2007) Hunter, J. D. 2007, Computing in Science & Engineering, 9, 90, doi: 10.1109/mcse.2007.55
  • Jayawardhana & Ivanov (2006) Jayawardhana, R., & Ivanov, V. D. 2006, Science, 313, 1279, doi: 10.1126/science.1132128
  • Joergens (2008) Joergens, V. 2008, A&A, 492, 545, doi: 10.1051/0004-6361:200810413
  • Kraus et al. (2014) Kraus, A. L., Ireland, M. J., Cieza, L. A., et al. 2014, ApJ, 781, 20, doi: 10.1088/0004-637x/781/1/20
  • Lazzoni et al. (2023) Lazzoni, C., Rice, K., Zurlo, A., Hinkley, S., & Desidera, S. 2023, MNRAS, 527, 3837, doi: 10.1093/mnras/stad3443
  • Lissauer (1993) Lissauer, J. J. 1993, ARA&A, 31, 129, doi: 10.1146/annurev.aa.31.090193.001021
  • Luhman et al. (2007) Luhman, K. L., Allers, K. N., Jaffe, D. T., et al. 2007, ApJ, 659, 1629, doi: 10.1086/512539
  • Martín et al. (2024) Martín, E. L., Žerjal, M., Bouy, H., et al. 2024, arXiv, doi: 10.48550/arxiv.2405.13497
  • McCaughrean & Pearson (2023) McCaughrean, M. J., & Pearson, S. G. 2023, arXiv, doi: 10.48550/arxiv.2310.03552
  • McCaughrean & Stauffer (1994) McCaughrean, M. J., & Stauffer, J. R. 1994, AJ, 108, 1382, doi: 10.1086/117160
  • Meylan & Mayor (1986) Meylan, G., & Mayor, M. 1986, A&A, 166, 122
  • Morales-Calderón et al. (2011) Morales-Calderón, M., Stauffer, J. R., Hillenbrand, L. A., et al. 2011, ApJ, 733, 50, doi: 10.1088/0004-637x/733/1/50
  • Mróz et al. (2017) Mróz, P., Udalski, A., Skowron, J., et al. 2017, Nature, 548, 183, doi: 10.1038/nature23276
  • Muench et al. (2008) Muench, A., Getman, K., Hillenbrand, L., & Preibisch, T. 2008, Handbook of Star Forming Regions, Volume I: The Northern Sky, Vol. 4, Star Formation in the Orion Nebula I: Stellar Content, ed. B. Reipurth (ASP Monograph Publications), doi: 10.48550/arxiv.0812.1323
  • Nesvorný & Morbidelli (2012) Nesvorný, D., & Morbidelli, A. 2012, AJ, 144, 117, doi: 10.1088/0004-6256/144/4/117
  • Ochiai et al. (2014) Ochiai, H., Nagasawa, M., & Ida, S. 2014, ApJ, 790, 92, doi: 10.1088/0004-637x/790/2/92
  • Pearson & McCaughrean (2023) Pearson, S. G., & McCaughrean, M. J. 2023, arXiv, doi: 10.48550/arxiv.2310.01231
  • Perryman et al. (1998) Perryman, M. A. C., Brown, A. G. A., Lebreton, Y., et al. 1998, A&A, 331, 81, doi: 10.48550/arxiv.astro-ph/9707253
  • Pollack et al. (1996) Pollack, J. B., Hubickyj, O., Bodenheimer, P., et al. 1996, Icarus, 124, 62, doi: 10.1006/icar.1996.0190
  • Portegies Zwart & Hochart (2023) Portegies Zwart, S., & Hochart, E. 2023, arXiv, doi: 10.48550/arxiv.2312.04645
  • Portegies Zwart et al. (2010) Portegies Zwart, S. F., McMillan, S. L., & Gieles, M. 2010, ARA&A, 48, 431, doi: 10.1146/annurev-astro-081309-130834
  • Prosser et al. (1994) Prosser, C. F., Stauffer, J. R., Hartmann, L., et al. 1994, ApJ, 421, 517, doi: 10.1086/173668
  • Rein & Liu (2012) Rein, H., & Liu, S.-F. 2012, A&A, 537, A128, doi: 10.1051/0004-6361/201118085
  • Rein & Spiegel (2014) Rein, H., & Spiegel, D. S. 2014, MNRAS, 446, 1424, doi: 10.1093/mnras/stu2164
  • Spurzem et al. (2009) Spurzem, R., Giersz, M., Heggie, D. C., & Lin, D. N. C. 2009, ApJ, 697, 458, doi: 10.1088/0004-637x/697/1/458
  • Tremaine (2023) Tremaine, S. 2023, Dynamics of Planetary Systems (New Jersey: Princeton University Press)
  • Wang et al. (2024) Wang, Y., Perna, R., & Zhu, Z. 2024, Nature Astronomy, 1, doi: 10.1038/s41550-024-02239-2
  • Yu & Lai (2024) Yu, F., & Lai, D. 2024, arXiv, doi: 10.48550/arxiv.2403.07224