Hot Jupiter and Ultra-cold Saturn Formation in Dense Star Clusters

, , , and

Published 2020 December 22 © 2020. The American Astronomical Society. All rights reserved.
, , Citation Yi-Han Wang et al 2020 ApJ 905 136 DOI 10.3847/1538-4357/abc619

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

0004-637X/905/2/136

Abstract

The discovery of high incidence of hot Jupiters in dense clusters challenges the field-based hot Jupiter formation theory. In dense clusters, interactions between planetary systems and flyby stars are relatively common. This has a significant impact on planetary systems, dominating hot Jupiter formation. In this paper, we perform high precision, few-body simulations of stellar flybys and subsequent planet migration in clusters. A large parameter space exploration demonstrates that close flybys that change the architecture of the planetary system can activate high eccentricity migration mechanisms: LK and planet–planet scattering, leading to high hot Jupiter formation rate in dense clusters. Our simulations predict that many of the hot Jupiters are accompanied by "ultra-cold Saturns," expelled to apastra of thousands of astronomical units. This increase is particularly remarkable for planetary systems originally hosting two giant planets with semimajor axis ratios of ∼4 and the flyby star approaching nearly perpendicular to the planetary orbital plane. The estimated lower limit to the hot Jupiter formation rate of a virialized cluster is $\sim 1.6\times {10}^{-4}{(\sigma /1\mathrm{km}{{\rm{s}}}^{-1})}^{5}{({a}_{{\rm{p}}}/20\mathrm{au})({M}_{{\rm{c}}}/1000{M}_{\odot })}^{-2}$ Gyr−1 per star, where σ is the cluster velocity dispersion, ap is the size of the planetary system, and Mc is the mass of the cluster. Our simulations yield a hot Jupiter abundance that is ∼50 times smaller than that observed in the old open cluster M67. We expect that interactions involving binary stars, as well as a third or more giant planets, will close the discrepancy.

Export citation and abstract BibTeX RIS

1. Introduction

The first exoplanet discovered around a main-sequence star, 51 Peg b (Mayor & Queloz 1995), with a very tight orbit, challenged astronomers' general picture of planetary system formation. This gas giant orbits about 10 times closer to its host star than similar planets in our solar system, which are only found outwards of 5 au. The discovery of the first so-called hot Jupiter not only initiated a revolution in planetary formation science, but also opened up the field of exoplanet discovery and characterization to include systems not observed in our own solar system. Because they are among the easier extrasolar planets to observe, hot Jupiters have been rapidly discovered in the past decade, adding to the large sample of exoplanets. The increasingly large number of hot Jupiters indicates that they play a major role in the field of exoplanets. As such, they have inspired extensive theoretical work on exoplanet formation, thereby broadening our understanding of planetary system formation and evolution.

Three main hot Jupiter formation theories have been proposed in the past decades. These are in situ formation, disk migration, and high eccentricity tidal migration. In the in situ formation theory, a hot Jupiter can be formed if a gravitational instability creates bound clumps from the fragments of the protoplanetary disk (Boss 1997; Durisen et al. 2007), or if the protoplanetary core accretes large quantities of gas from the protoplanetary disk (Perri & Cameron 1974; Pollack et al. 1996; Chabrier et al. 2014). Batygin et al. (2016) have argued that hot Jupiters can form with super-Earth cores in the inner parts of disks. Bailey & Batygin (2018) have argued that the inner boundary of the period-mass distribution of close-in giant planets conforms to the expectations of the in situ formation scenario. A recent strong argument against in situ formation is the measurement (Gunther & Guinan 2020) of a period (and hence semimajor axis (SMA) decrease of the hot Jupiters in HD 189733 and six other systems.

In the gas disk migration theory, torques from the gaseous protoplanetary disk may significantly shrink a giant planet's orbit to form a hot Jupiter (Goldreich & Tremaine 1980; Lin & Papaloizou 1986; Lin et al. 1996; Ida & Lin 2008; Brucalassi et al. 2014). The efficiency of disk migration depends on the strength of the corotation torques which is in turn determined by the disk conditions. If the migration timescale is shorter than the disk lifetime, the giant may migrate all the way into the Roche radius and be tidally disrupted by its host star. If the migration timescale is comparable to or greater than the disk lifetime, the giant planet may not migrate far enough inwards to make a hot Jupiter.

In high eccentricity tidal migration, several dynamical mechanisms have been proposed to excite the eccentricity of a cold Jupiter to high values. This is done by removing/exchanging angular momentum with the giant planet's orbit, so that tidal migration becomes efficient enough to act on sufficiently short timescales to make a hot Jupiter. The proposed mechanisms are planet–planet scattering that converts Keplerian sheer into an angular momentum deficit and triggers high eccentricity migration (Rasio & Ford 1996; Weidenschilling & Marzari 1996; Ford & Rasio 2006; Chatterjee et al. 2008), secular interactions, including LK oscillations from the companion star (Wu & Murray 2003; Fabrycky & Tremaine 2007; Katz et al. 2011; Naoz et al. 2012; Storch et al. 2014; Storch & Lai 2014; Petrovich 2015; Anderson et al. 2016), LK oscillations from another giant planet (Naoz et al. 2011; Teyssandier et al. 2013), secular chaos from multiple planets (Wu & Lithwick 2011; Xu & Lai 2016; Hamers et al. 2017; Li et al. 2019, 2020), and coplanar secular effects (Li et al. 2014).

No single mechanism mentioned above has yet been able to explain the distribution of orbital properties of the observed hot Jupiters. Ford et al. (2003) and Ford & Rasio (2006) proposed that multiple planet scatterings (i.e., sequential scatterings acting on the same planet but separated in time) may be promising to explain the observations, but this ultimately fails due to a low efficiency in producing hot planets. Adams (2005) demonstrated that the combination of stellar scattering and subsequent tidal migration may produce the observed hot Jupiter properties. This illustrates the possibility of tweaking some of the existing theories for producing hot Jupiters by introducing more degrees of complexity. However, this is nontrivial since further complications arise from the fact that a large fraction of stars are born in clusters or associations. Stellar interactions (scatterings) within a cluster/association can significantly change the orbits of the planets or the disk (Hurley & Shara 2002; Adams et al. 2006; Spurzem et al. 2009; Li & Adams 2015; Shara et al. 2016). These perturbed planetary systems may eventually escape into the field from their birth environment due to the dissociation of the cluster.

Indeed, planets have been detected in star clusters (Lovis & Mayor 2007; Sato et al. 2007). Meibom et al. (2013) detected two small planets in the open cluster NGC 6811, suggesting that planets are as common in open clusters as in the field. Quinn et al. (2012) and Brucalassi et al. (2016) reported the discovery of four giant planets orbiting M67 stars, including three hot Jupiters. Quinn et al. (2014) reported an eccentric hot Jupiter in the Hyades open cluster. All of these observations provide promising evidence that stellar interactions in clusters/associations may affect the planet formation process and/or perturb their orbital parameters post-formation on short timescales (i.e., shorter than the timescale for cluster disruption), and thus may play a nontrivial role in hot Jupiter formation.

Hurley & Shara (2002), Spurzem et al. (2009), Parker & Quanz (2012), and Hamers & Tremaine (2017) showed that the orbits of single planets in a cluster can be significantly affected by stellar encounters by performing large-scale N-body simulations. Shara et al. (2016) and Li et al. (2020) have shown that scatterings due to passing stars in a star cluster can change the orbits of two-planet planetary systems, producing both hot Jupiters and very distant Saturns. Stellar encounters can clearly induce external perturbations that significantly change the architecture of multiple planet systems in clusters/associations, and can thus affect the subsequent internal dynamics within the planetary system (see also recent works by Cai et al. 2019; Flammini Dotti et al. 2019; Li et al. 2020; Wang et al. 2020a, 2020b). We naively predict that multi-planet systems that begin close to orbital resonances will be less susceptible to changes to their orbital configurations due to stellar interactions (i.e., flybys). This is because (small) integer ratios in the orbital periods mark stabilized subsets of the parameter space. Many mechanisms can act to stabilize orbital resonances (see Murray & Dermott 1999 for a more detailed discussion). One important example is due to fine-tuning in the degree of the imparted impulse at specific points along the orbits. Strong perturbations are needed to knock multi-planet systems well away from preexisting resonances. Hence, if a system is perturbed only weakly away from an orbital resonance, it is very likely to secularly migrate back to this same orbital resonance (or another nearby resonance). Conversely, if multi-planet systems do not begin with any such strong resonances, we naively expect to see stronger variations in the final distributions of orbital parameters, relative to cases where one or more initial resonances are present. In this paper, we perform scattering experiments both close to and far away from strong resonances (i.e., with both large and small ratios between orbital periods), in order to study and quantify these competing effects.

How do stellar encounters alter the efficiency of hot Jupiter formation? To answer this question in full, a sophisticated simulation is needed which includes cluster evolution, stellar evolution, planet formation, single star/binary star scatterings, and effects related to the internal dynamical evolution of planetary systems, such as tidal migration and/or planet–planet scatterings. In this paper, we take a first step in this direction by performing scatterings between single star planetary systems hosting two giant planets, Jupiter-like and Saturn-like. After the scattering, tidal migration models are adopted to track the formation channels of the hot Jupiters with full few-body simulations. Figure 1 shows our schematic illustration of the stellar flyby and tidal migration process.

The paper is organized as follows. In Section 2 we describe the numerical models we use. Modifications of planetary systems' architectures due to flybys are described in Section 3. In Section 4 we describe the formation of hot Jupiters from high eccentricity tidal migration, including LK and planet–planet scattering effects. We discuss our results and summarize our conclusions in Section 5.

2. Numerical Method

The scattering experiments are performed using the high precision few-body code SpaceHub (details in Wang et al. 2018, 2019). We use the scattering facilities of SpaceHub to generate the initial conditions. The following subsections present the initial conditions and the assumptions going into our numerical scattering experiments.

2.1. Flyby Process

2.1.1. Planetary System Setups

The planetary system consists of a one solar mass main-sequence star and two giant planets. The inner planet has a mass equal to Jupiter's, and its orbit is circular with a 5 au radius. The outer planet has the mass of Saturn, in a circular orbit in the same plane as the inner planet. We explore three different SMA values for the outer planet: 10, 20, and 40 au. The mean anomalies (i.e., the fraction of an orbit's period that has elapsed since the orbiting body passed periapsis) of the two planets are generated uniformly between 0 and 2π, independently. The uniform distribution of mean anomalies ensures that the phase of the planet is uniformly distributed in time.

2.1.2. Critical Velocity for Multi-hierarchy Scattering

The critical velocity vc is an important variable in multi-body scattering. It is defined such that, for a relative velocity at infinity between two objects v > vc (objects can be isolated bodies or they can have complex internal orbital architectures as is the case for binaries, triples, etc.), the total energy of the system in the center of mass reference frame is positive (e.g., Hut & Bahcall 1983; Fregeau et al. 2004; Leigh & Geller 2012; Leigh et al. 2016a, 2016b). Thus, in the regime where the total interaction energy is positive, full ionization is a possible outcome scenario for objects containing internal orbital architectures. In this regime, the incident object usually does a flyby relative to the scattered object, and interacts with it only once. For v < vc, the total energy of the scattering system is negative. The corresponding scatterings are likely to be resonant in this regime, such that a bound chaotic interaction ensues, with the incident object undergoing repeated flybys relative to the scattered object. These interactions typically progress until disruption, forming at least one tightly bound system (other orbits need not survive to satisfy the criteria for energy and angular momentum conservation).

We label the masses of the k components of the scattered system to be m00, m01, ⋯ m0k and the masses of the l components of the incident system to be m10, m11, ⋯ m1l . If the relative velocity between the centers of mass of the two objects is v, the total energy of the scattering system in the center of mass reference frame is

Equation (1)

where M0 = ∑i m0i is the total mass of the scattered object and M1 = ∑i m1i is the total mass of the incident object. V0 and V1 are the initial center of mass velocities of the scattered and incident objects, respectively. The critical velocity between two objects corresponding to a total energy of zero for the system is calculated using the scattering facilities of SpaceHub, specifically via the following equation:

Equation (2)

2.1.3. Impact Parameter for High Mass Ratio Scattering

Due to gravitational focusing, incident objects at spatial infinity with impact parameter b will pass the scattered object at closest approach with a corresponding critical pericenter distance for the hyperbolic/parabolic orbit:

Equation (3)

We only consider interactions satisfying Q < Qc, where pc is the critical pericenter distance. Therefore, to cover all parameter space in which the desired outcome is causally and physically permitted, the scattering experiments must be sampled with an impact parameter distribution of b2, uniformly distributed in the range $[0,{b}_{\max }^{2}]$, where bmax is the corresponding impact parameter for pc. We can rewrite Equation (3) in the more convenient form:

Equation (4)

For higher values of v, ${b}_{\max }\to {Q}_{{\rm{c}}}$, while the lower limit gives ${b}_{\max }\sim \tfrac{1}{{v}_{\infty }}$. In Hut & Bahcall (1983), the two limits are combined to construct the following equation:

Equation (5)

where R is the contact distance between the two scattering objects (e.g., single-binary scattering is the SMA of the binary; binary–binary scattering is the sum of the semimajor axes of the binaries) and C and D are dimensionless constants that depend on the desired outcome (DR ∼ Qc in the high v limit and CR ∼ Qc in the low v limit). This equation is widely adopted in nearly equal mass single-binary and binary–binary scatterings. However, in this work, the planet mass is much smaller than the star mass, and hence special consideration should be given to this interesting limit. The velocity term v in the original equation is made dimensionless by the critical velocity vc. If we take the limit of high v in both Equations (4) and (5), we have

Equation (6)

since CR ∼ Qc. The equation above holds only if

Equation (7)

Equation (2) then requires

Equation (8)

For single-binary scatterings with component masses (m00, m01)–m1, this becomes

Equation (9)

If m00 ∼ m01 ∼ m1, this yields C ∼ 4, which is the frequently adopted value. However, if m01 is a Jupiter mass planet, or m00 ∼ m1 ∼ 1000m01, we obtain C ∼ 2000. Indeed,

Equation (10)

Therefore, for high mass ratio systems, we cannot use Equation (5) by normalizing v by vc and letting C ∼ [4–10] (Hut & Bahcall 1983; Fregeau et al. 2004). The scattering facilities of SpaceHub provide both choices given in Equations (4) and (5).

2.1.4. Tidal Factor Between Two Objects

The tidal factor is widely used in scattering experiments to estimate the tidal perturbation between two objects. It is defined as

Equation (11)

where Ftid,0→1 is the tidal force that object 0 exerts on object 1 and Frel,1 is the internal gravitational force of object 1. If δ → 0, then object 0 and object 1 are isolated. This tidal tolerance factor is often used to save on computational runtime, by switching to integrating Keplerian orbits in the limit of very low tidal tolerance factors (i.e., this option is turned on whenever the tidal tolerance parameter drops below a critical user-defined value). In this paper, we only use this parameter to generate the initial position of the interloper but we do not invoke this approximation afterwards, and explicitly integrate the N-body problem without any such simplifying approximations.

In binary–binary and binary-single scatterings, as in Fregeau et al. (2004), we have

Equation (12)

Equation (13)

Indeed, a more general equation for scatterings between any two generic objects is provided by the tidal radius

Equation (14)

where D1 is the geometric size of object 1 (i.e., the stellar radius for a single star, the orbital separation for a binary star, etc.). For binary-single scatterings we have D1 = a1(1 + e1). The typical value δ ∼ 0.02 yields the classical tidal disruption equation

Equation (15)

The scattering facilities of SpaceHub take the maximum value of either rtid,0→1(δ0→1) or rtid,1→0(δ1→0) as the final tidal radius rtid(δ).

For scattering experiments, strong interactions occur if p < pc ∼ rtid(δ = 0.02), while perturbative interactions occur if rtid(δ = 0.02) < p < rtid(δ = 10−4). The tidal radius is also used to calculate the initial relative distance between the scattered and incident objects. In this project, we take rtid(δ = 10−5) as the initial distance separating the centers of mass for the two scattering objects. The trajectories from infinity to this initial distance are calculated analytically by regarding the scattering objects as isolated up until this critical point.

2.1.5. Initial Conditions for Orbital Angles

The phases of the scattered and incident objects are isothermally distributed. For each orbit, cos(i) is uniformly distributed in the range [−1, 1], Ω is uniformly distributed in the range [−π, π], ω is uniformly distributed in the range [−π, π], and M is uniformly distributed in the range [−π, π], where i is the orbital inclination, Ω is the longitude of the ascending node, ω is the argument of periapsis, and M is the mean anomaly.

2.1.6. Termination Time of Flyby

The interloper is dropped at the point where the distance between the center of mass of the interloper and the center of mass of the planetary system is rtid(δ = 10−5). In the host cluster environment, we adopt the 1D velocity dispersions σ ∼ 0.2 and 1 km s−1, following Shara et al. (2016), which are typical of open star clusters (e.g., Geller & Leigh 2015; Leigh et al. 2016b). The corresponding relative velocities at infinity are, respectively, v ∼ 0.73 and ∼3.6 km s−1. The velocity is not in the resonant interaction regime, i.e., v > vc. Therefore, we terminate the flyby process at t = 2tdrop, where the drop-in time tdrop is the time that the interloper takes to travel from the starting point to the point of closest approach. In the hyperbolic asymptotic limit,

Equation (16)

where M is the relative mean anomaly between the interloper dropping point and the distance of closest approach, a is the corresponding SMA of the incident hyperbolic orbit and Mtot is the total mass of the system.

2.2. Tidal Migration Process

After the flyby, the planetary system is perturbed, and high eccentricity excitations can occur due to several channels, including planet–planet scatterings (Rasio & Ford 1996; Weidenschilling & Marzari 1996; Ford & Rasio 2006; Chatterjee et al. 2008), LK oscillations (Wu & Murray 2003; Fabrycky & Tremaine 2007; Katz et al. 2011; Naoz et al. 2012, 2011; Teyssandier et al. 2013; Petrovich 2015), and coplanar secular interactions (Li et al. 2014). The (excited) high eccentricity effectively increases the rate of tidal dissipation due to the tidal force at pericenter, which makes planet migration possible. We integrate the after-flyby planetary systems up to 1 Gyr, and discuss the results of the flybys at 0.01, 0.1, and 1 Gyr.

2.2.1. Tidal Dissipation

We adopt the static tidal mode by implementing the tidal perturbation acceleration between stellar particles and planet particles in the simulation. The tidal acceleration that the star exerts on the planet is

Equation (17)

where r, Rp, kAM,p and τp are the particle distance, the planetary radius, the apsidal motion constant and the tidal time lag, respectively. Ωp is the planetary spin frequency, where we use the orbit-averaged assumption (Hut 1981; Hamers et al. 2017),

Equation (18)

where $n=\sqrt{{{GM}}_{* }/{a}^{3}}$ is the mean motion of the planetary orbit and a and e are the SMA and the eccentricity, respectively. In this work, we use kAM,p = 0.25 and τp = 0.66 s (Hamers et al. 2017). Typically, the timescale for stellar tides to operate is much longer than for planetary tides. Therefore, we do not take it into consideration.

In the N-body code, the variables a and e are calculated from the instant relative positions and velocities between particles, that is,

Equation (19)

Equation (20)

where μ = GMtot is the gravitational parameter of the planet–star pair and d r and d v are the relative position and velocity between the planet–star pair, respectively. Since the accelerations are calculated between particle pairs in the code, the variables d r and d v are always calculated between particles. However, for systems with more than two particles, once a hierarchical triple system is formed, the outer object will orbit around the center of mass of the inner binary. Therefore, to calculate a and e of the outer orbit, the d r and d v must be the relative position and velocity between the outer object and the center of mass of the inner binary. In our case, the inner binary is a star–planet system, and hence the center of mass of the inner binary is close to the star. Therefore, it is a reasonably good approximation to simply use position and velocity of the star to calculate a and e of the outer orbit.

2.2.2. General Relativistic (GR) Effects

In secular eccentricity excitation processes (including the standard LK and coplanar secular processes), the periapsis of the inner orbit precesses around its center of mass. However, the general relativity effect precesses the orbit in the opposite direction of the LK oscillations, and can thus significantly suppress the eccentricity excitation (Liu et al. 2015). In this work, we implement GR precession via the first order post-Newtonian (PN) method. The corresponding pair-wise acceleration is

Equation (21)

The second order PN term contributes very little to the GR precession, which can be regarded as a correction to the first order PN term with the efficiency being lower by an order of magnitude in v/c. Thus, it can be safely ignored in our calculations. The 2.5 PN gravitational wave radiation term in the planetary system can also be ignored due to its very long timescale. The timescale for merger due to gravitational radiation in a circular isolated binary with initial SMA a0 is (Peters 1964)

Equation (22)

where m12 is the total mass of the isolated binary and MJ is Jupiter mass. For an eccentric isolated binary, the corresponding timescale is

Equation (23)

In order for the timescale τGW(e ∼ 1) to be smaller than the age of the universe, e needs to be larger than ∼0.999, which happens in only a tiny fraction of our simulations even when eccentricity excitation mechanisms are important.

3. Flyby Modifications of Planetary Systems

Flybys can significantly change the architecture of the target planetary system such that the modified system may undergo tidal migration due to several different mechanisms. While the single interloper flyby problem and the mechanisms that lead to high eccentricity have previously been studied (e.g., Hamers & Samsing 2019a, 2019b), it is instructive to see the fraction of hot Jupiters that emerge via different formation channels in interacting stellar environments. These effects are quantified below in the following sections.

We mainly consider planet–planet LK oscillations, planet–planet scattering, and coplanar secular interactions. In order to operate, all of these mechanisms require that the planetary system retains two planets. However, the flyby star may cause ejections of one or both planets, especially when the impact parameter is small. Clearly we must first answer the question: what is the fraction of systems that retain two planets immediately after the flyby? How does that fraction vary with impact parameter, incident angles, and planet SMA ratio?

To answer these questions, we perform groups of flyby scatterings with as/aj = [2, 4, 8], i = [0°, 30°, 60°, 90°], Q = [1/8, 1/4, 1/2, 1, 2, 4, 8] as, and σ = [1, 0.2] km s−1, where as is the SMA of the outer (Saturn) planet, aj is the SMA of the inner (Jupiter) planet, i is the incident inclination of the interloper, Q is the closest approach distance between the interloper and the planetary system, and σ is the velocity dispersion of the stellar cluster. In each setup, prograde and retrograde interlopers are uniformly generated in equal numbers.

The outcomes immediately after the flyby are classified into seven sets:

Jupiter ejection: The inner planet is ejected due to the stellar flyby, but the outer planet remains bound to its original host star.

Saturn ejection: The outer planet is ejected due to the stellar flyby, but the inner planet remains bound to its original host star.

Both ejected: Both planets are ejected due to the stellar flyby.

Both remain: Both planets remain bound to the original host star.

Star–star collisions: Collisions between two stars when the radii of the objects overlap.

Star–planet collisions: Collisions between a star and a planet when the radii of the objects overlap.

Planet–planet collisions: Collisions between two planets when the radii of the objects overlap.

Figure 1.

Figure 1. Schematic illustration of the flyby and the various outcomes after long timescale interactions. After the flyby, the planet system is modified. If the system remains intact, with both planets still orbiting their host star, the modified triple may form a new configuration in the planet–planet (p-p) interaction regime, the LK regime and the coplanar secular regime. In those regimes, the eccentricity of the Jupiter can be excited to high values, which makes hot Jupiter formation possible due to tidal circularization with GR effects.

Standard image High-resolution image

3.1. Orbital and System Architectures After the Flyby for Different Velocity Dispersions

We have performed more than 1 million scattering experiments with fixed incident angles, closest approach distances, SMA ratios, and cluster velocity dispersions as described above. The orbits of all planets that remain bound to their host stars, described by their eccentricity—SMA distributions, are shown in Figures 25.

Figure 2.

Figure 2. The distribution of orbital parameters of the planets that remain bound at four different times: upper left panel, immediately after the flyby; upper right panel, 106 yr after the flyby; bottom left panel, 107 yr after the flyby; and bottom right panel, 108 yr after the flyby. The velocity dispersion is σ = 1 km s−1 and the initial SMA of the Jupiter and Saturn are 5 and 10 au, respectively.

Standard image High-resolution image
Figure 3.

Figure 3. The distribution of orbital parameters of the planets which remain bound at four different times: upper left panel, immediately after the flyby; upper right panel, 106 yr after the flyby; bottom left panel, 107 yr after the flyby; and bottom right panel, 108 yr after the flyby. The velocity dispersion is σ = 1 km s−1 and the initial SMA of the Jupiter and Saturn are 5 and 20 au, respectively.

Standard image High-resolution image
Figure 4.

Figure 4. The distribution of orbital parameters of the planets that remain bound at four different times: upper left panel, immediately after the flyby; upper right panel, 106 yr after the flyby; bottom left panel, 107 yr after the flyby; and bottom right panel, 108 yr after the flyby. The velocity dispersion is σ = 1 km s−1 and the initial SMA of the Jupiter and Saturn are 5 and 40 au, respectively.

Standard image High-resolution image
Figure 5.

Figure 5. The distribution of orbital parameters of the planets which remain bound at four different times: upper left panel, immediately after the flyby; upper right panel, 106 yr after the flyby; bottom left panel, 107 yr after the flyby; and bottom right panel, 108 yr after the flyby. The velocity dispersion is σ = 0.2 km s−1 and the initial SMA of the Jupiter and Saturn are 5 and 10 au, respectively.

Standard image High-resolution image

Several features of Figures 25 encapsulate some of the key findings of this paper. They illustrate the time evolution of the orbits of Jupiters and Saturns that survive an encounter, for the cases as/aj = 2, 4, and 8. The numbers of Jupiters reaching Saturn's orbit, and vice versa, is largest for as/aj = 2. Perhaps most striking is that hot Jupiters are never immediately formed as an aftermath of a close flyby. In contrast, both warm Jupiters (with aj ∼ 1 au) and "ultra-cold" Saturns (with as ∼ 100–10,000 au are immediately formed, and they persist for longer than 100 Myr. By 106 yr hot Jupiters have been formed, and increase in number at 107 and 108 yr due to tidal circularization. Ultra-cold Saturns decrease in number as these weakly bound objects occasionally acquire enough energy to escape their host stars.

The different orbital outcomes for a range of incident angles and closest approaches by the intruder star, immediately after the flyby, and at 106, 107, and 108 yr, respectively, are shown in Figures 69. Both Saturn and Jupiter's initial orbits are negligibly perturbed when the intruder's closest approach is further than twice Saturn's SMA, and these are omitted from the figures.

Figure 6.

Figure 6. Distribution of orbital parameters right after the flyby. The velocity dispersion is σ = 1 km s−1 and the initial SMA of the Jupiter and Saturn are 5 and 20 au, respectively.

Standard image High-resolution image
Figure 7.

Figure 7. Distribution of orbital parameters at 106 yr after the flyby. The velocity dispersion is σ = 1 km s−1 and the initial SMA of the Jupiter and Saturn are 5 and 20 au, respectively.

Standard image High-resolution image
Figure 8.

Figure 8. Distribution of orbital parameters at 107 yr after the flyby. The velocity dispersion is σ = 1 km s−1 and the initial SMA of the Jupiter and Saturn are 5 and 20 au, respectively.

Standard image High-resolution image
Figure 9.

Figure 9. Distribution of orbital parameters at 108 yr after the flyby. The velocity dispersion is σ = 1 km s−1 and the initial SMA of the Jupiter and Saturn are 5 and 20 au, respectively.

Standard image High-resolution image

As Figure 6 shows, a flyby star approaching as close as Saturn strongly perturbs that planet on the flyby timescale. It sends some Saturns inwards as close to their host stars as Jupiter, and flings others outwards as far as 10,000 au, beginning their ultra-cold Saturn existence. Strongly preferred and avoided regions in a–e phase space are obvious in this figure. On timescales of 106–108 yr both planets migrate enough, under their mutual gravitational influence and tides, to smear out the sharp a–e features of Figure 6, as shown in Figures 79.

Figure 7 shows that hot Jupiters are produced within 106 yr of the encounter. Figure 7 suggests, and Figures 8 and 9 confirm, that high inclination flybys with closest approaches inside Jupiter's orbit are by far the most effective way to make hot Jupiters.

In each subplot of Figures 69, with fixed i and Q, the initial phases of the two planets are uniformly generated in time with eccentricity equal to zero. We see that for perturbations with Q > 2ap, where ap is the SMA of the outer planet (Saturn), the flyby star hardly changes the SMA of the Jupiter, regardless of the incident inclinations. With Q < 2ap and even Q < ap, where the interloper penetrates the planetary system, there are no directly formed hot Jupiters with ${a}_{{\rm{j}}}^{{\prime} }\lt 0.09\,\mathrm{au}$. This is because the flyby star must extract most of the energy from the Jupiter orbit to make it shrink below ∼0.09 au, which is almost impossible within a single flyby. From this plot we also see that, for a given i and Q, the cluster velocity dispersion has little effect on the distribution of the SMA of the Jupiter right after the flyby. This indicates that, in the strong interaction regime Q < ap, the energy change of the inner planet is insensitive to the velocity dispersion of the environment. This is because the velocity of the interloper at the closest approach is

Equation (24)

where Mtot is the total mass of the scattering system. For a velocity dispersion σ = 0.2 or 1 km s−1, v is negligible compared to 2GMtot/Q with small Q. Therefore, the term vQ is not sensitive to the environmental velocity dispersion in the strong interaction regime.

In each subplot, we also see clear differences between the prograde and retrograde orbits. For scatterings with fixed Q but different i, we see that in the prograde scatterings, the Jupiter shrinks its SMA more than in the case of retrograde scatterings. Since in the prograde scattering the interloper has more time to interact with the Jupiter at the point of closest approach due to the low relative velocity between the interloper and the planets, the interloper is able to extract more energy from Jupiter's orbit. This trend begins to disappear as the inclination i increases to 90°. In the 90° case, with the interloper traveling perpendicular to the planet orbital plane, the interaction time between the interloper and the planet is the same in the prograde and retrograde scatterings.

As shown in each row of Figure 10, as i increases, the SMA changes become smaller, due to the shorter interaction time between the interloper and planet at the point of closest approach. As i increases, the maximum energy that can be extracted from the planetary system becomes smaller.

Figure 10.

Figure 10. The distribution of the SMA of the Jupiter right after the flyby. The green vertical line shows the initial distance. The initial SMA ratio as/aj = 2 is fixed with aj = 5 au. Each subplot shows the result for fixed incident inclination angle i and distance of closest approach Q.

Standard image High-resolution image

We also see in many of the subplots in Figure 10 that there are two peaks in both prograde and retrograde scatterings. In total, we have four peaks from different relative positions at closest approach: (i) prograde scattering with the interloper and planet at the same side of the star (the interloper and planet move in the same direction with small relative distance, thus contributing to the dominant peak), (ii) retrograde scattering with the interloper and planet at the opposite side of the star (the interloper and planet move in the same direction with large relative distance), which contributes to the secondary dominant peak, (iii) prograde scattering with the interloper and planet at opposite sides of the star (the interloper and planet move in opposite directions with large relative distance), which contribute to the third prominent peak, and (iv) retrograde scattering with the interloper and planet at the same side of the star (the interloper and planet move in the opposite direction with small relative distance), which contributes to the weakest peak that is almost invisible.

The distance between the peaks becomes smaller as Q increases. The reason is that, for large Q, the distance between the interloper and planet, Q ± aj, is dominated by Q instead of by ±aj. Thus, no matter which side of the star the planet is located on, i.e., +aj or −aj, the distance between the interloper and the planet in either case becomes similar.

Figure 11 shows the distribution of the SMA of the Saturn right after the flyby. Because the outer planet orbit is more weakly bound, it is easier to modify the orbit of the Saturn relative to that of the Jupiter. In general, the orbit of the Saturn right after the flyby shows properties similar to those of the Jupiter. However, we see more differences arising when adopting different velocity dispersions, especially in prograde scatterings. This is because, with fixed Q, the relative velocity between the interloper and the planets determines the interaction time between the interloper and the planet, which affects the properties of the planet right after the flyby. The relative velocity between the interloper and planet at the distance of closest approach is determined by the planet's orbital velocity and the velocity of the interloper at closest approach vQ. In the case of Jupiter, due to the tight orbit, the orbital velocity is dominant over the velocity of the interloper, thus the velocity dispersion does not affect much the relative velocity between the interloper and the planet at the point of closest approach. However, for the Saturn with smaller orbital velocity, the orbital velocity becomes less dominant over the velocity of the interloper. Thus, the velocity dispersion has a larger impact on the relative velocity between the interloper and the Saturn at the point of closest approach.

Figure 11.

Figure 11. The distribution of the SMA of the Saturn right after the flyby. The initial conditions are the same as in Figure 10.

Standard image High-resolution image

The SMA change of the planets can be estimated by

Equation (25)

where ap is the initial SMA and ${a}_{{\rm{p}}}^{{\prime} }$ is the SMA immediately after the flyby. It is straightforward to note that, if (Ep − Ep,0)/∣Ep,0∣ < 0, the corresponding orbit shrinks, while if 0 < (Ep − Ep,0)/∣Ep,0∣ < 1, the orbit expands. For (Ep − Ep,0)/∣Ep,0∣ > 1, the planet is ejected from its host star. As the incident angle increases, the maximum relative binding energy shift becomes smaller as the interaction time at the point of closest approach decreases. Therefore, the binding energy shift becomes weaker as Q increases.

The prograde scatterings are responsible for the majority of the planet ejections, where (Ep − Ep,0)/∣Ep,0∣ > 1. The prograde scatterings with the interloper and planets at the same side of the host star have the longest interaction times (and hence impart the largest impulse) needed to remove the binding energy from the planet. Since the retrograde encounter is not as efficient as the prograde case in removing the binding energy of the planet due to its shorter interaction time, the retrograde scattering can hardly change the SMA of the Jupiter at Q ∼ 4aj and the SMA of the Saturn at Q ∼ 4as, whereas prograde scatterings do change the SMA of the planet significantly (by up to about one order of magnitude).

Figures 12 and 13 show the eccentricity of the Jupiter and Saturn right after the flyby. Since the initial value of the eccentricity of the planets is zero, this also indicates the eccentricity change right after the first flyby. For Q < 4aj and Q < 4as (i.e., the strong interaction regime), a flyby can significantly kick the eccentricity of a planet to a large value. For Q < aj or Q < as, where the interloper penetrates within the orbit of the planetary system, the majority of the planets are perturbed into the "near ejection" state where e ∼ 1. As for Q > 4aj and Q > 4as, where scattering happens in the perturbative regime, the eccentricity shift due to the flyby rapidly drops below 0.1. These results indicate that just one flyby does not efficiently increase the eccentricity of a planet, which is essential to enable the subsequent high eccentricity tidal migration process.

Figure 12.

Figure 12. The distribution of the eccentricity of the Jupiter immediately after the flyby. The initial conditions are the same as in Figure 10.

Standard image High-resolution image
Figure 13.

Figure 13. The distribution of the eccentricity of the Saturn right after the flyby. The initial conditions are the same as in Figure 10.

Standard image High-resolution image

From Figures 12 and 13, we can also see that as the inclination i increases, retrograde scatterings become similar to the prograde scatterings in removing the angular momentum of the planets.

Figure 14 shows the inclination between the Jupiter and Saturn orbits right after the flyby if both the Jupiter and Saturn still remain bound to their original host star. When the closest approach Q is smaller than ap, we see that it is possible for the interloper to flip the orbital orientation of the Saturn to the opposite direction by transferring sufficient angular momentum to its orbit. Also, as the incident inclination i increases, we see that the probability of creating an inclined Saturn orbit with respect to the Jupiter orbit increases as well, especially in the prograde flyby case. It is difficult to get a highly inclined orbit when Q > ap, such that LK oscillations are unlikely to operate on the system after the flyby to help make hot Jupiters.

Figure 14.

Figure 14. The distribution of the inclination between the two planet orbits right after the flyby. The initial conditions are the same as in Figure 10.

Standard image High-resolution image

In Figures 1014, we explore the properties of the planetary system right after the flyby with two different values of the environmental velocity dispersion, σ = 0.2 or 1 km s−1. We find that close flybys create an ideal situation for tidal migration to produce hot Jupiters. Close flybys can excite the eccentricity of the planets by extracting significant angular momentum from planetary systems and kick the original coplanar planetary system into an inclined planetary system. These two processes increase the efficiency of the tidal migration process.

We find that the orbital properties are insensitive to the low velocity dispersion. The velocity of the interloper at infinity contributes very little to the velocity of the interloper at the distance of closest approach for small Q due to the strong gravitational focusing, as indicated in Equation (24).

3.2. System Architecture After the Flyby with Different Initial SMA Ratios

Besides exploring how the cluster velocity dispersion changes the orbital properties of the planetary system after the flyby, we also explore the flyby scattering with different initial SMA ratios. We fix the velocity dispersion to be 1 km s−1 and perform scatterings with initial SMA ratios equal to 2, 4, and 8.

Before proceeding, we briefly comment on the possible impact of orbital resonances on our results. The naive expectation here is that, for smaller initial SMA ratios, such resonances should be more important as the initial SMA ratio is closer to a strong resonance (e.g., the 2:1 resonance for an initial SMA ratio of 2). As the initial ratio of SMAs increases, we expect resonances to become less important, as these resonances would tend to operate at very high ratios of the orbital periods, where we expect resonances to be very weak and have little to no dynamical effects. In our simulations, we see only mild evidence for systems becoming locked into orbital resonances for initial SMA ratios of 2, with at most of order a few percent of our total simulations ending up stabilized in a resonant state. For higher initial SMA ratios, these effects disappear. This is expected since the perturbed planetary systems will be scattered closest to mainly higher-order resonances, which tend to get weaker as the ratio of periods increases. In the end, we find that orbital resonances have a very negligible effect on our overall results, statistics, and conclusions, at least for the initial conditions considered here. Thus, these effects can safely be ignored in this section when trying to understand how the product distributions from our simulations depend on the initial SMA ratio.

As in Figure 10, Figure 15 shows the SMA of the Jupiter right after the flyby for different initial planet SMA ratios (see also Figures 1619). It is indicated in the figure that, as we double the SMA ratio, the distribution of aj is almost identical to the distribution of aj with the original ratio but with Q doubled, e.g., the as/aj = 8 with Q = 0.125ap is almost identical to as/aj = 4 with Q = 0.25ap. It is not surprising that the absolute closest approach determines the post-flyby SMA and eccentricity (as shown in Figure 17) of the Jupiter. The reason is that the inter-planet interaction in the close flyby is negligible due to the short interaction flyby time. Thus, we can treat the perturbation exerted on each planet almost independently.

Figure 15.

Figure 15. The distribution of the SMA of the Jupiter right after the flyby, for different initial SMAs of the Saturn. The velocity dispersion is σ = 1 km s−1 and the initial SMA of Jupiter is aj = 5 au. Each subplot shows the result for fixed incident inclination angle i and distance of closest approach Q.

Standard image High-resolution image
Figure 16.

Figure 16. The distribution of the SMA of the Saturn right after the flyby. The initial conditions for σ, aj, and the three values of as are the same as in Figure 15.

Standard image High-resolution image
Figure 17.

Figure 17. The distribution of the eccentricity of the Jupiter immediately after the flyby. The initial conditions are the same as in Figure 15.

Standard image High-resolution image

As already noted and discussed, orbital inclinations are created by highly inclined flybys. The changing z-direction to the interloper changes the direction and magnitude of the angular momentum of the planets. For fixed Q, as the SMA ratio decreases, the SMA of the Jupiter decreases while keeping the SMA of the Saturn constant. Since the angular momentum of the planet is

Equation (26)

then, for a smaller SMA, it is easier to change the direction of the inner planet by importing/extracting the same angular momentum to/from the planet's orbit. Hence, we can see in Figure 19 that, with a smaller SMA ratio, the planetary system tends to be more inclined after the flyby. This argument is also valid in coplanar flybys: smaller SMA ratios also create more retrograde orbits in all those interactions corresponding to the closest approaches.

Figure 18.

Figure 18. The distribution of the eccentricity of the Saturn right after the flyby. The initial conditions are the same as in Figure 15.

Standard image High-resolution image
Figure 19.

Figure 19. The distribution of the inclination between the two planet orbits right after the flyby. The initial conditions are the same as in Figure 15.

Standard image High-resolution image

We can further analyze the SMA ratio after the flyby with the same method by treating the perturbation exerted on each planet independently. The binding energy of the planet is

Equation (27)

Therefore, it is harder to change the SMA of the planet with smaller initial SMA by importing/extracting the same energy to/from the planet's orbit. With fixed Q, a smaller initial SMA ratio gives a smaller absolute SMA for the Jupiter. It becomes more difficult to change the SMA of the Jupiter. Thus, the final SMA ratio right after the flyby depends more on the change in SMA of the Saturn.

Since the absolute SMA of the planets determines if the interaction between the planet and the interloper is in the fast/slow regime, it is difficult to make a simple argument on how the final SMA ratio depends only on the initial SMA ratio. For instance, if the initial SMA ratio is 4, where both Saturn and Jupiter reside in the slow regime where the flyby tends to increase the SMA, then the final SMA tends to become larger. The reason is that, while the SMA of both planets tends to become larger, the change in Saturn's SMA is more significant. If the initial SMA ratio is 2, Jupiter's orbit may enter into the faster regime, where the flyby tends to tighten Jupiter's orbit while keeping the Saturn in the slow regime. Thus, although smaller initial SMA ratios make the final SMA ratio more weakly dependent on Jupiter's orbit, the final SMA ratio tends to become larger than in the case where the initial SMA ratio is equal to 4 because the two planets migrate in opposite directions.

To make the planets interact more efficiently, the SMA ratio of the planets right after the flyby should not be too large nor too small. Two close orbits make planet–planet scattering possible during the subsequent long-term evolution after the flyby and shorten the timescale for secular processes to operate (under the condition that the system is stable).

Because of the short flyby time, the interaction between the planets during the flyby at the point of closest approach can be safely ignored. Hence, we can treat the perturbation exerted by the interloper on the two planets independently. Therefore, the scattering with different initial SMA ratios affects the properties of the planets independently, as shown in Figures 1518.

However, the initial SMA ratio affects the inclination and final SMA of the planetary system right after the flyby, thus influencing the probability and timescale for subsequent tidal migration. We will discuss the corresponding "sweet spot" for the initial SMA ratio in the next section.

3.3. Event Fraction Right After the Flyby

As listed in Section 3, the outcomes after the flyby are classified into "planet–planet collision," "star–planet collision," "star–star collision," "Jupiter ejection," "Saturn ejection," "both ejection," and "both remain." In this subsection we will discuss the fraction of each outcome for different environmental velocity dispersions and initial SMA ratios. We pay special attention to the both-remain systems since these are obviously required for post-interaction tidal migration.

Figure 20 shows the standard case where as/aj = 2 and σ = 1 km s−1. We can see that for Q > 4ap, the flyby will never change the architecture of the planetary system, and therefore all the outcomes are both remain. As the closest approach comes close to the planetary system with Q ∼ 2ap, the flyby begins to eject the outer planet Saturn and then eject the inner planet Jupiter. As shown in the row with Q = 2ap, high inclination incident angles where the interaction time is shorter decrease the ejection fraction for Saturn alone.

Figure 20.

Figure 20. The event fraction right after the flyby. Each subplot shows the result for fixed incident inclination i and closest approach Q. The initial SMA ratio is as/aj = 2 with aj = 5 au and the velocity dispersion σ is 1 km s−1.

Standard image High-resolution image

The fraction of the both-remain systems in the close flyby regime where Q < ap is roughly 10%. We also find that most of the collisions are planet–star collisions that only appear in coplanar scatterings. The reason for this is related to the volume-filling factor of the objects, which is much smaller for coplanar configurations relative to the isotropic case (Leigh et al. 2017, 2018). In other words, there is more free space in the isotropic case, increasing the timescale for two objects to coincide in both time and space.

As an aside we note that planet–star collisions can give rise to a rich set of possible outcomes (De Marco & Soker 2011; Schuler et al. 2011; Ginsburg et al. 2012; Kratter & Perets 2012; Veras et al. 2013; Deal et al. 2015). In particular, planets that are engulfed may shape the ejecta of their host stars, while swallowed and disrupted planets may enrich the outer layers of stars in detectable lithium, calcium, magnesium, and iron.

Figure 21 shows the results for the same initial conditions as in Figure 20, but with σ equal to 0.2 km s−1. As discussed before, v contributes only minimally to the velocity of the interloper at the distance of closet approach vQ. Therefore, the fractions are very similar to the standard case with σ equal to 1 km s−1. Hence, we may conclude that in stellar clusters with low velocity dispersions where gravitational focusing is significant, the fraction of planetary systems that may form hot Jupiters after a close flyby is not particularly sensitive to the velocity dispersion.

Figure 21.

Figure 21. Same as in Figure 20 but with a velocity dispersion σ of 0.2 km s−1.

Standard image High-resolution image

Figures 22 and 23 show the cases where as/aj = 4 and 8, respectively. With a larger SMA ratio, we see a larger fraction in the ratio of Saturn ejections, and a much lower both-ejection fraction. Since planet ejection requires a close encounter with the interloper, it is more difficult for the interloper to have a close encounter with both planets if the two planets are far apart.

Figure 22.

Figure 22. Same as in Figure 20 but with an initial SMA ratio as/aj = 4.

Standard image High-resolution image
Figure 23.

Figure 23. Same as in Figure 20 but with an initial SMA ratio as/aj = 8.

Standard image High-resolution image

4. Hot Jupiters from High eccentricity Tidal Migration

Flybys alone cannot directly create hot Jupiters, as this would require the flyby star to extract large amounts of angular momentum without draining much of the orbital energy of the planet. However, flybys can significantly perturb the architecture of the planetary system into regimes where several mechanisms can then excite the eccentricity to very high values that enable efficient tidal dissipation. Hot Jupiters can then be formed from high eccentricity tidal migration. Isolated planet tidal dissipation can be described by (e.g., Murray & Dermott 1999; Hamers et al. 2018; Samsing et al. 2018),

Equation (28)

Equation (29)

where

Equation (30)

The two equations converge to ${a}_{{\rm{p}}}\to {a}_{{\rm{p}},0}(1-{e}_{{\rm{p}},0}^{2})$ and ep → 0 as t → +  . Due to the $(1-{e}_{{\rm{p}}}^{2})$ term in the denominator, the dissipation rate increases rapidly as ep,0 → 0. High eccentricities are clearly critical in hot Jupiter formation via dynamical processes. Several mechanisms can excite the eccentricity of the planet to high values, as discussed below.

4.1. Lidov–Kozai (LK) Oscillations

In a hierarchical triple system, the inner orbit (here, the star-Jupiter binary) can undergo LK oscillations. In LK oscillations, the energy (∝ − Mtot,in/a) of each orbit is conserved but the inner and outer orbits continuously exchange angular momentum ($\propto \sqrt{a(1-{e}^{2})}$). Therefore, the eccentricity of the inner orbit can be excited to very high values as its angular momentum reaches a minimum, which is balanced to conserve angular momentum by reducing the orbital inclination between the inner and outer orbits. The timescale for the quadrupole LK cycle to operate is

Equation (31)

where nin is the mean motion of the Jupiter orbit, Mtot, in is the total mass of the inner orbit, and Mout is the mass of the perturber. LK oscillations require the triple system to be hierarchical and stable. These oscillations operate in the range where $-\sqrt{3/5}\lt \cos (I)\lt \sqrt{3/5}$, with I being the inclination between the two planets' orbits.

In this work, the flyby perturbation generates the necessary inclination, and the two giant planet orbits ensure that the timescale τLK is shorter than the age of the universe. This combination of factors allows the eccentricity of the Jupiter to be increased to a high enough value that tidal dissipation becomes efficient enough to make a hot Jupiter.

An example of hot Jupiter formation from the LK effect is shown in Figure 24. It can be seen that the effect starts to operate on a timescale of about 1–1.5 Myr.

Figure 24.

Figure 24. An example of hot Jupiter formation from the LK effect. This is clearly visible as operating most effectively between 1 and 1.5 Myr, after which time Jupiter's orbit circularizes.

Standard image High-resolution image

4.2. Planet–Planet Scattering

Close planet–planet scatterings can convert Keplerian shear into an angular momentum deficit, triggering high eccentricity migration (e.g., Rasio & Ford 1996; Weidenschilling & Marzari 1996; Ford & Rasio 2006; Chatterjee et al. 2008). In planet–planet scattering, the planets typically undergo several close encounters to grow their eccentricities to high values. However, if the orbital velocity of a planet is larger than the escape velocity at the planet's surface, the cross section for collision would be larger than the cross section for scattering. This may lead to a planet collision rather than just eccentricity growth. Goldreich et al. (2004), Ida et al. (2013) and Petrovich et al. (2014) give the upper limit of the eccentricity from planet–planet scattering as

Equation (32)

where Mp is the mass of planet, Rp is the radius of the planet, ap is the SMA of the planet, and P is the orbital period of the planet. A Jupiter mass planet with ap ∼ 1 au has roughly ep−p ∼ 1, which makes high eccentricity tidal migration possible.

An example of hot Jupiter formation via planet–planet scattering is shown in Figure 25. This occurs rather continuously up to a time of about 13 Myr, when the orbit of Jupiter achieves circularization. After that time, Jupiter and Saturn no longer interact strongly.

Figure 25.

Figure 25. An example of hot Jupiter formation from planet–planet scattering. This operates relatively continuously up until about 13 Myr, after which time Jupiter's orbit circularizes, terminating the strong interactions between Jupiter and Saturn.

Standard image High-resolution image

4.3. Coplanar Secular Octupole LK Effect

LK oscillations operate only when the orbital inclination is within a certain range. However, Li et al. (2014) found that a perturber with a highly elliptical orbit, where the octupole LK effect is strong, can also excite the eccentricity of an initially coplanar system. Thus, tidal migration may also occur in this regime. Indeed, we find that usually the flyby star extracts the angular momentum of the outer planet rather than that of the inner planet. This preferentially forms an elliptical outer orbit for the Saturn, which drives the further evolution of the system. The coplanar octupole effect comes into play when

Equation (33)

where ein is the initial eccentricity of the inner binary, ω is the argument of the periapsis of the outer orbit, and epsilon is the LK octupole parameter

Equation (34)

No hot Jupiters were found to be formed via this channel in our simulations.

4.4. Other Mechanisms

There are other mechanisms that can excite the eccentricity of a planet, including LK oscillations from a stellar companion (Naoz et al. 2012; Anderson et al. 2016) and secular chaos due to multiple LK timescales. Stellar LK oscillations demand a companion star, either from binary star formation as an outcome of the flyby (which is very unlikely) or a flyby binary interloper, which is beyond the scope of this work. Secular chaos generally demands more than two planets, also beyond the scope of this paper. Thus, we do not consider eccentricity excitation from these mechanisms.

4.5. Hot Jupiter Formation Fraction from Different Channels

After the flyby, the modified planetary system may fall into the eccentricity excitation regime, where hot Jupiter formation is possible. For all the both-remain systems, as discussed in more detail in the preceding section, we divide the outcomes into several different regimes based on the criteria below:

LK: The inclination between the two planets' orbits satisfies $-\sqrt{3/5}\lt \cos (I)\lt \sqrt{3/5}$ and the quadrupole timescale is τLK < 1 Gyr.

Planet–planet scattering: The pericenter of the outer planet/the apocenter of the inner planet satisfies $\tfrac{{a}_{{\rm{s}}}(1-{e}_{{\rm{s}}})}{{a}_{{\rm{j}}}(1+{e}_{{\rm{j}}})}\lt 1$, where the planets could undergo close encounters.

Coplanar octupole: The inclination between the two planet orbits is within 5° and Equation (33) is satisfied.

4.6. Long-term Jupiter Evolution After the Flyby

After the flyby, we perform long-time integrations of the perturbed planetary systems up to 1 Gyr, including tidal dissipation and GR effects as described in Section 2.2. Since for Q > 2ap the perturbation becomes too weak to transition the planetary system into the high eccentricity tidal migration regime, we only perform long integrations for systems with Q < 2ap.

During the long-time evolution, tidal dissipation and GR effects may break the planetary system by ejecting either the Jupiter or Saturn through a close planet–planet encounter or collision. If the eccentricity of the remaining planet is not high enough to make tidal dissipation efficient, then we will not observe it to form a hot Jupiter within the age of the universe.

Figure 26 shows the distribution of the disruption time (i.e., the timescale over which the planetary system is disrupted) in log scale. For the coplanar case, this timescale is much shorter than for the other cases. The planetary system tends to disrupt in ∼104 yr, while the other cases are more likely disrupt at around 106 yr. This is because for the coplanar case, there is a much higher probability for the two planets to undergo close encounters. The distribution is truncated by the flyby time on the left side and by the integration upper limit on the right side.

Figure 26.

Figure 26. The distribution of the planetary system disruption time in the long interaction time regime. The velocity dispersion of the cluster is 1 km s−1, the initial SMA of the Jupiter is 5 au, and the initial SMA of the Saturn is 10 au.

Standard image High-resolution image

Figure 27 shows the SMA of the Jupiter right after the flyby (gray), 106 yr after the flyby (blue), 107 yr after the flyby (purple), and 108 yr after the flyby (orange). We see that, for Q < 0.5ap, a non-negligible fraction of the Jupiters evolve into hot Jupiters with aj between 10−2 and 10−1 au due to tidal dissipation and the aforementioned high eccentricity excitation mechanisms. For Q > 1 ap, it is extremely difficult to form a hot Jupiter progenitor. We do not see any hot Jupiters even 108 yr after the flyby.

Figure 27.

Figure 27. Snapshots of the distribution of the SMA of the Jupiter right after the flyby, after 106 yr of evolution, after 107 yr of evolution, and after 108 yr of evolution. The velocity dispersion of the cluster is 1 km s−1, the initial SMA of the Jupiter is 5 au and the initial SMA of the Saturn is 10 au.

Standard image High-resolution image

The evolution of the orbital SMA of Jupiter with time, at several representative times up to 1 Gyr, is shown in Figure 28. For most configurations, hot Jupiter formation happens promptly, already present at 1 Myr.

Figure 28.

Figure 28. Evolution of the orbital SMA of Jupiter with time, shown at some representative times up to 1 Gyr.

Standard image High-resolution image

Figure 29 shows the same distribution as Figure 27 but with an initial SMA ratio equal to 4. A large fraction of the Jupiters evolve into hot Jupiters with Q = 0.125ap and i = 90°. This is the ideal parameter space where the flyby has a large probability to create a highly inclined system with an appropriate SMA ratio that boosts the LK effect. For an initial SMA ratio equal to 2, as indicated in Figure 27, the two giant planets are too close to maintain the architecture of the planetary system. Thus, there is a higher chance for the planetary system to undergo close planet–planet interactions rather than secular LK excitations. For an initial SMA ratio equal to 8, as shown in Figure 30, we see that the fraction of the hot Jupiters drops again in highly inclined flybys. The timescale of the quadrupole LK effect is given by Equation (31). Unlike the LK mechanism in binary star systems, where the perturber is a star, the perturber in the planetary system is only a giant planet. Thus, the system requires a much smaller SMA ratio to make the LK effect operate within the age of the cluster. Therefore, with the larger SMA ratio equal to 8, the LK effect becomes less efficient than for an SMA ratio equal to 4. For a Jupiter mass planet-solar mass host star binary with aj = 5 au and with a Saturn mass perturber at 10 au, the typical timescale given by Equation (31) that is sensitive to the SMA ratio is ∼2.5 × 106 yr. With an initial SMA ratio equal to 8, the timescale is about one order of magnitude longer. This explains why the hot Jupiter formation fraction drops quickly for aj/as = 8. However, the SMA ratio cannot be too small in the LK regime, otherwise it breaks the stability of the planetary system. Therefore, for aj/as = 2, the main source of hot Jupiters is from planet–planet scatterings.

Figure 29.

Figure 29. Snapshots of the distribution of the SMA of the Jupiter right after the flyby, after 106, 107, and 108 yr of evolution. The velocity dispersion of the cluster is 1 km s−1, the initial SMA of the Jupiter is 5 au, and the initial SMA of the Saturn is 20 au.

Standard image High-resolution image
Figure 30.

Figure 30. Snapshots of the distribution of the SMA of the Jupiter right after the flyby, after 106, 107, and 108 yr of evolution. The velocity dispersion of the cluster is 1 km s−1, the initial SMA of the Jupiter is 5 au, and the initial SMA of the Saturn is 40 au.

Standard image High-resolution image

Figures 3133 show the distribution of the eccentricity of the Jupiter right after the flyby (gray), 106 yr after the flyby (blue), 107 yr after the flyby (purple), and 108 yr after the flyby (orange). In the close flyby regime where Q < 2aj, the general trend of the Jupiter eccentricity after long interaction times is to decrease, while in the perturbative regime where Q > 2aj the general trend of the Jupiter eccentricity is to increase. Indeed, in the close flyby regime, the interloper increases the eccentricity of the Jupiter by extracting angular momentum from the planet. There is a large probability of creating a Jupiter orbit with a small pericenter. However, in the perturbative regime, the flyby interloper barely creates a small pericenter Jupiter orbit. Therefore, in the close flyby regime, the tidal dissipation will dominate in the long-time integration, while in the perturbative regime, planet–planet interactions become the dominant effect. The tidal dissipation process will circularize the Jupiter orbit but planet–planet interactions will increase the eccentricity of the Jupiter orbit. Therefore, we see opposite general trends for the eccentricity evolution in the close flyby and perturbative regimes.

Figure 31.

Figure 31. Snapshots of the distribution of the eccentricity of the Jupiter right after the flyby, after 106, 107, and 108 yr of evolution. The velocity dispersion of the cluster is 1 km s−1, the initial SMA of the Jupiter is 5 au, and the initial SMA of the Saturn is 10 au. The initial eccentricities are zero.

Standard image High-resolution image
Figure 32.

Figure 32. Snapshots of the distribution of the eccentricity of the Jupiter right after the flyby, after 106, 107, and 108 yr of evolution. The velocity dispersion of the cluster is 1 km s−1, the initial SMA of the Jupiter is 5 au, and the initial SMA of the Saturn is 20 au. The initial eccentricities are zero.

Standard image High-resolution image
Figure 33.

Figure 33. Snapshots of the distribution of the eccentricity of the Jupiter right after the flyby, after 106, 107, and 108 yr of evolution. The velocity dispersion of the cluster is 1 km s−1, the initial SMA of the Jupiter is 5 au, and the initial SMA of the Saturn is 40 au. The initial eccentricities are zero.

Standard image High-resolution image

Figure 34 shows the hot Jupiter formation fraction of all both-remain systems after the flyby. Different colors represent the fractions from different channels and the sum of the fractions indicates the total formation fraction. We find that there is no hot Jupiter formed from the coplanar secular effect due to its very narrow parameter space, such that flybys cannot create the ideal initial conditions in this regime. For as/aj = 2, most of the hot Jupiters formed from planet–planet interactions. For as/aj = 4, the LK channel contributes most of the hot Jupiters, especially the perpendicular flybys with a formation fraction one order of magnitude higher. For as/aj = 8, we see that most of the hot Jupiters formed from the LK channel. Indeed, with as/aj = 2, the flybys more easily create closer planet orbits to activate planet–planet scatterings. However, with as/aj = 8, where the two orbits are distant from each other, it is harder for the flybys to bring the two orbits close enough together. Therefore, we see that hot Jupiter formation from planet–planet scattering dominates in the small SMA ratio regime, while hot Jupiter formation from secular LK effects dominates in the large SMA ratio regime. Figure 35 shows the formation fraction with lower velocity dispersion, where σ = 0.2  km s−1. The formation fraction is slightly higher than for σ = 1 km s−1.

Figure 34.

Figure 34. Hot Jupiter formation fraction of different channels with different initial SMA ratios at 108 yr after the flyby. The velocity dispersion of the cluster is 1 km s−1, the initial SMA of the Jupiter is 5 au. For Q > ap, the fraction is smaller than 0.1%, and no hot Jupiter formed from the coplanar secular channel.

Standard image High-resolution image
Figure 35.

Figure 35. Hot Jupiter formation fraction for different channels with different velocity dispersions at 108 yr after the flyby. The initial SMA of the Jupiter is 5 au and the initial SMA of the Saturn is 10 au. For Q > ap, the fraction is smaller than 0.1% and no hot Jupiter formed from the coplanar secular channel.

Standard image High-resolution image
Figure 36.

Figure 36. Rate of close flybys as a function of the velocity dispersion in the host virialized cluster, and for different values of the cluster mass. The dashed lines indicate the cluster number densities.

Standard image High-resolution image

4.7. Hot Jupiter Formation Cross Section and Rate

To estimate the hot Jupiter formation rate from different channels in different environments, we need the hot Jupiter formation fractions of each channel from isotropic scatterings. Then, the hot Jupiter formation rate in each channel can be estimated by

Equation (35)

where ${\sigma }_{{{\rm{a}}}_{p}}$ is the corresponding cross section for Q = 1ap for which a hot Jupiter can be formed (Q < ap), n is the number density of the environment, vrms is the root mean square velocity of the environment, and f is the hot Jupiter formation fraction of each channel after a close flyby. The cross section is given by

Equation (36)

where

Equation (37)

The number density of a virialized cluster is (e.g., Binney & Tremaine 1987; Spitzer 1987)

Equation (38)

where Mc and $\bar{m}$ are the total stellar mass and mean stellar mass of the cluster, respectively. The upper panel of Figure 36 shows the number density of the virialized open, young massive and globular clusters as a function of the velocity dispersion and the bottom panel of Figure 36 shows the number of the "close" flyby per star as a function of the velocity dispersion. Then, the hot Jupiter formation rate can be estimated from

Equation (39)

with ${M}_{0}\sim {M}_{1}\sim \bar{m}$, ${v}_{\mathrm{rms}}^{2}=2{v}_{\infty }^{2}$, and $G\bar{m}/{a}_{{\rm{p}}}\sim {v}_{{\rm{p}}}^{2}$. For most clusters vrms ≪ vp, therefore,

Equation (40)

Since the hot Jupiter formation timescale is typically much shorter than the age of its host star cluster, the hot Jupiter percentage per star in a cluster is

Equation (41)

where τc is the age of the cluster.

We performed isotropic scatterings with cluster velocity dispersions σ =  0.2 and 1 km s−1. The impact parameter b of each scattering was randomly generated from the probability distribution function f(b) ∝ b within [0, bmax], where bmax is chosen to be large enough that the system is not significantly perturbed. Based on our scattering experiments described in the last section, the planetary systems after the flyby perturbation will be almost unchanged if Q > ap. Therefore, we set the corresponding b with Q = ap as the bmax in isotropic scatterings. All other setups are the same as described in Section 2.

Finally, it is worth emphasizing that the properties of star clusters change over time, primarily due to two-body relaxation and the evaporation of the cluster in a Galactic tidal field. We have not accounted for these effects in any systematic way in this paper, and instead defer properly accounting for these effects to future work. Briefly, the rate of stellar escape is inversely proportional to stellar mass, causing older clusters to become more depleted in low-mass stars and tending toward a top-heavy stellar mass function (relative to standard initial mass functions measured in young star-forming regions). In future work, a Monte Carlo approach will be needed for a given host galaxy to generate a realistic initial star cluster mass function, a realistic distribution of 3D star cluster galactocentric distances, and so on.

5. Summary and Conclusions

In this work we have studied hot Jupiter and ultra-cold Saturn formation in an environment of strong stellar interactions. Close flybys in stellar clusters can significantly change the configurations of the planetary systems that host the progenitors of hot Jupiters, significantly affecting the rate of hot Jupiter formation. While primarily focused on the hot Jupiters, our simulations yield the full reconfiguration of the planetary system, hence allowing us to track also the distribution of orbital parameters of the outer planet, the Saturn, after the flyby.

By performing high precision few-body scatterings in the flyby process and long-term integrations with tidal dissipation and GR corrections up to 1 Gyr, our main conclusions can be summarized as follows:

  • 1.  
    In stellar clusters where the velocity dispersion is relatively low, close encounters (where closest approach is approximately the size of the planetary system) that involve strong interactions give similar post-scattered configurations. In low velocity dispersion environments, the relative velocity at closest approach that determines the imparted impulse are mostly contributed by gravitational focusing, resulting in a weak dependence on the velocity dispersion.
  • 2.  
    Without additional secular effects via additional degrees of freedom, it is improbable to create a hot Jupiter directly from a single close flyby via the extraction of a large amount of angular momentum from the planetary system. Hot Jupiter formation in dense clusters demands long-term interactions, including perturbations from other bodies and subsequent tidal dissipation.
  • 3.  
    Close flybys can activate high eccentricity tidal migration of a Jupiter by changing the SMA ratio of the planets and the eccentricity of the outer planet. Close flybys in dense clusters thus significantly increase the hot Jupiter formation rate.
  • 4.  
    Distant flybys, where the closest approach is much larger than the size of the planetary system (Q ≫ ap), do not by themselves activate high eccentricity tidal migration of Jupiters. They thus contribute little to the hot Jupiter formation rate in dense clusters.
  • 5.  
    About 10% of close flybys can activate high eccentricity tidal migration by significantly changing the configuration of the planetary system. The remaining 90% of the outcomes correspond to planet ejections, making star–planet/star–star/planet–planet collisions rare.
  • 6.  
    The most common formation channels that are activated by flybys are the LK mechanism (i.e., the outer giant planet acts as the perturber) and planet–planet scatterings. With small initial planetary SMA ratios, planet–planet scattering is the main mechanism operating to create hot Jupiters. For larger initial ratios, LK activation is the main mechanism creating hot Jupiters.
  • 7.  
    It is extremely rare to create hot Jupiters from coplanar secular effects in dense clusters. Thus, this naively predicts that retrograde hot Jupiters born in dense clusters should be rare.
  • 8.  
    Perpendicular close flybys more frequently activate hot Jupiter formation from LK oscillations for moderately small initial SMA ratios (2  ≲ as/aj  ≲ 4). Therefore, for even larger initial SMA ratios (i.e., as/aj ≳4), perpendicular close flybys become even less efficient at activating hot Jupiter formation from LK oscillations.
  • 9.  
    Most hot Jupiters in dense clusters are generated within 108 yr of the encounter that triggered them. We find that the distribution of hot Jupiters converges within 109 yr of the triggering encounter.
  • 10.  
    The dynamical reconfiguration produced by the flybys leads to frequent occurrences in which the Saturn moves onto a very wide orbit, larger than ∼100 au in many systems, but frequently exceeding ∼1000 au. Hence, flybys can readily explain the observed presence of giant planets on very wide orbits (ultra-cold Saturns) (Lafrenière et al. 2011), which are not easily explained by conventional theories of planet formation.
  • 11.  
    The properties of star clusters change over time, primarily due to two-body relaxation and the evaporation of the cluster in a Galactic tidal field. We have not accounted for these effects in any systematic way in this paper, and instead defer properly accounting for these effects to future work, in which we perform a more sophisticated population synthesis-based study accounting for a realistic star cluster mass function, distribution of galactocentric distances, and so on.

Professors Ward Wheeler and Cheryl Hayashi of the Richard Guilder Graduate school at the American Museum of Natural History made generous grants of CPU cycles on their computation clusters available to us, which we gratefully acknowledge. N.W.C.L. gratefully acknowledges the support of a Fondecyt Iniciacion grant #11180005.

Please wait… references are loading.
10.3847/1538-4357/abc619