Tidal Evolution of the Eccentric Moon around Dwarf Planet (225088) Gonggong

, , , and

Published 2021 November 3 © 2021. The American Astronomical Society. All rights reserved.
, , Citation Sota Arakawa et al 2021 AJ 162 226 DOI 10.3847/1538-3881/ac1f91

Download Article PDF
DownloadArticle ePub

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

1538-3881/162/6/226

Abstract

Recent astronomical observations revealed that (225088) Gonggong, a 1000 km sized trans-Neptunian dwarf planet, hosts an eccentric satellite, Xiangliu, with an eccentricity of approximately 0.3. As the majority of known satellite systems around trans-Neptunian dwarf planets have circular orbits, the observed eccentricity of the Gonggong–Xiangliu system may reflect the singular properties of the system. In this study, we assumed that the Gonggong–Xiangliu system formed via a giant impact and we investigated the subsequent secular tidal evolution of this system under the simplifying assumptions of homogeneous bodies and of zero orbital inclination. We conducted simulations of coupled thermal–orbital evolution using the Andrade viscoelastic model and included higher-order eccentricity functions. The distribution of the final eccentricity from a large number of simulations with different initial conditions revealed that the radius of Xiangliu is not larger than 100 km. We also derived the analytical solution of the semilatus rectum evolution, a function of the radius of Xiangliu. From the point of view of the final semilatus rectum, the radius of Xiangliu was estimated to be close to 100 km. Together with the results of the Hubble Space Telescope observations, our findings suggest Gonggong and Xiangliu have similar albedos.

Export citation and abstract BibTeX RIS

1. Introduction

There are six (or seven) known trans-Neptunian objects (TNOs) with diameters larger than 1000 km, and (almost) all of them host one or more satellites orbiting the primary (e.g., Brown et al. 2006; Parker et al. 2016; Kiss et al. 2017). 4 The discovery of satellites around large TNOs provided us with a key to understanding the early history of the outer solar system.

The orbital period of a satellite system allows the determination of its total mass. We can obtain the density of the primary body when the size is known from radiometry and/or occultation measurements and estimate the bulk composition (e.g., the ice-to-rock ratio; Brown 2012). Other orbital elements, such as the eccentricity and inclination, are also important for understanding how satellites form and evolve. The majority of moons around 1000 km sized TNOs have circular orbits. For example, the orbit of the Pluto–Charon system is nearly circular (e ∼ 5 × 10−5, where e is the eccentricity; Brozović et al. 2015). Based on the spatially resolved observations from the Atacama Large Millimeter/submillimeter Array by Brown & Butler (2018), the eccentricity of the Eris–Dysnomia system is also small (e < 4 × 10−3). Their circular orbits are thought to be the result of tidal evolution, because satellite systems are initially eccentric if they formed via giant impacts (e.g., Canup 2005; Arakawa et al. 2019).

However, the two known exceptions to this trend are the Quaoar–Weywot and Gonggong–Xiangliu systems. Using the adaptive optics facility of the Keck II telescope, Fraser et al. (2013) revealed that the orbit of the Quaoar–Weywot system is moderately eccentric (e = 0.13–0.16). Kiss et al. (2019) also reported an eccentricity of e ≃ 0.3 for the orbit of the Gonggong–Xiangliu system. If the tidal dissipation primarily occurs inside the satellite rather than inside the host dwarf planets, the eccentricity decreases (e.g., Ward & Canup 2006; Cheng et al. 2014; Arakawa et al. 2019). Therefore, the eccentricity evolution depends strongly on the rheological properties of the satellites and their hosts. The variation in the observed eccentricity among satellite systems of large TNOs may reflect differences in the rheological properties, which should be related to their formation process and thermal histories.

The primary, (225088) Gonggong, was discovered by Schwamb et al. (2009). The satellite of (225088) Gonggong, Xiangliu, was found on archival images of the WFC3 camera of the Hubble Space Telescope obtained in 2009 and 2010 by Kiss et al. (2017). This initial discovery was based only on observations at two epochs, and the orbit of the satellite could not be derived. Kiss et al. (2019) conducted recovery observations of Xiangliu taken with the WFC3 camera in 2017 and determined the orbital elements of the Gonggong–Xiangliu system.

Table 1 shows the orbital elements and fundamental parameters of the Gonggong–Xiangliu system. The radius of Gonggong, RG, was also estimated by Kiss et al. (2019) using thermophysical modeling. The spin period of Gonggong was obtained from the light-curve analysis by Pál et al. (2016): PG,obs = 22.4 hr or 44.8 hr. The radius and spin period of Xiangliu have not yet been determined. Assuming the same albedo as the primary, Kiss et al. (2017) obtained the radius of Xiangliu as RX = 120 km. Kiss et al. (2019) also discussed the radius of Xiangliu, and used simple estimates of the eccentricity damping timescale to propose that 18 km < RX < 50 km might be suitable to explain the observed nonzero eccentricity of the system. The estimated RX corresponds to a geometric albedo of 1.0–0.2, which is higher than the estimated albedo for the primary (0.15 ± 0.02; Kiss et al. 2019).

Table 1. Orbital Elements and Fundamental Properties of Gonggong–Xiangliu Assumed in this Study

SymbolPropertyValueUnitReferences
aobs Observed semimajor axis2.4 × 104 kmKiss et al. (2019)
eobs Observed eccentricity0.3Kiss et al. (2019)
PG,obs Observed spin period of Gonggong 22.4 or 44.8hPál et al. (2016)
PX,obs Observed spin period of Xiangliuh
Mtot Total mass of the system1.75 × 1021 kgKiss et al. (2019)
RG Radius of Gonggong600 kmKiss et al. (2019)
RX Radius of Xiangliu km

Note. Parameters for standard runs are indicated by bold face. Kiss et al. (2019) reported two orbital solutions. One is called the prograde solution with an orbital period of Porb = 25.22073 ± 0.000357 days, semimajor axis of aobs = 24,021 ± 202 km, and eccentricity of eobs = 0.2908 ± 0.0070. The other is called the retrograde solution with an orbital period of Porb = 25.22385 ± 0.000362 days, semimajor axis of aobs = 24,274 ± 193 km, and eccentricity of eobs = 0.2828 ± 0.0063 (see Table 2 of Kiss et al. 2019). Assuming a coplanar primary equator and satellite orbit and a spherical shape of the primary, the effective diameter of Gonggong is Deff ≡ 2RG = 1224 ± 55 km and 1238 ± 50 km for the prograde solution for the spin periods of PG,obs = 44.8 hr and 22.4 hr, respectively, and Deff = 1227 ± 56 km and 1241 ± 50 km for the retrograde solutions with PG,obs = 44.8 and 22.4 hr (see Table 3 of Kiss et al. 2019). Then we set aobs = 24,000 km, eobs = 0.3, and RG = 600 km for simplicity.

Download table as:  ASCIITypeset image

Multiple studies address the formation and tidal evolution of the Pluto–Charon system (e.g., Canup 2005; Cheng et al. 2014; Barr & Collins 2015; Desch 2015; Kenyon & Bromley 2019; Rozner et al. 2020; Renaud et al. 2021; Bagheri et al. 2021; Shimoni et al. 2022). The formation and tidal evolution of other systems have also been discussed in several studies (e.g., Greenberg & Barnes 2008; Ortiz et al. 2012). Arakawa et al. (2019) found that the general trend of the final eccentricity of satellites around 1000 km sized TNOs depends strongly on their thermal history, based on higher-order calculations of tidal evolution.

Renaud et al. (2021) studied the tidal evolution of the Pluto–Charon system and found that considering higher-order eccentricity terms was necessary to calculate the tidal evolution of an eccentric system (considering terms beyond e20 would be necessary for systems with eccentricities greater than 0.8). Renaud et al. (2021) also found that higher-order spin–orbit resonances exist not only for highly eccentric systems but also for systems with small eccentricity of e ≃ 0.1. In most prior studies (e.g., Cheng et al. 2014; Arakawa et al. 2019), the strength of tidal dissipation was estimated by using the traditional model of constant phase lag (Kaula 1964). However, the strength of tidal dissipation and the stability of the higher-order spin–orbit resonances depend strongly on the rheological properties (e.g., Walterová & Běhounková 2020). Thus, we need to use a more realistic rheology model and simultaneously calculate the orbital and thermal evolution.

In this study, we assumed that the Gonggong–Xiangliu system formed via a giant impact as an intact fragment and we calculated the subsequent secular tidal evolution of the system under the simplifying assumption of homogeneous bodies (e.g., Ojakangas & Stevenson 1986; Shoji et al. 2013). Our study is the first attempt to simulate the secular tidal evolution of the Gonggong–Xiangliu system in detail. We investigated the coupled thermal–orbital evolution using a realistic rheology model and included higher-order eccentricity terms.

The structure of this paper is as follows. In Section 2, we discuss the initial condition of the Gonggong–Xiangliu system formed via the giant impact. In Section 3, we briefly introduce the orbital and thermal evolution models. In Section 4, we present the statistics of the final state of the satellite systems. We show the distributions of the final spin/orbital properties of the system in Sections 5 and 6 and summarize them in Section 7.

2. Initial Condition of the System

In this study, we assume that the Gonggong–Xiangliu system was formed by a giant impact and discuss the orbital properties of the two bodies immediately after the impact. We re-evaluate the results of the giant impact simulations by Arakawa et al. (2019) in Section 2.1 and discuss the initial condition from the point of view of the conservation of the angular momentum in Section 2.2.

2.1. Initial Orbit and Spin of Satellites Formed via Giant Impacts

We define γX,GMX/MG as the secondary-to-primary mass ratio of the system formed after giant impacts: satellite systems with 10−3γX,G ≲ 10−1 can be directly formed via giant impacts as intact fragments (Canup 2005; Arakawa et al. 2019). We also define γi,tMimp/Mtar as the impactor-to-target mass ratio, where Mtar and Mimp are the masses of the target and impactor, respectively. Arakawa et al. (2019) performed giant impact simulations with three settings for the targets and impactors: γi,t = 0.5 with an ice composition (Set A), γi,t = 0.2 with an ice composition (Set B), and γi,t = 0.5 with a basalt composition (Set C). The total mass of the target and the impactor is Mtar + Mimp = 6 × 1021 kg in the simulations.

Figure 1(a) shows the distribution of the semilatus rectum from the results of the giant impact simulations (Arakawa et al. 2019). The semilatus rectum, porb, and the periapsis distance, qorb, are defined as follows (Murray & Dermott 1999):

Equation (1)

Equation (2)

where a is the semimajor axis and e is the eccentricity. Arakawa et al. (2019) performed more than 400 runs of giant impact simulations of 1000 km sized TNOs, and revealed that the initial periapsis distance of satellites formed via giant impacts, qini, is in the range

Equation (3)

where RG is the radius of the primary. As the initial eccentricity of satellites formed via giant impacts, eini, is widely distributed in the range 0–1 (Arakawa et al. 2019), the initial semilatus rectum, pini, should be in the range

Equation (4)

We confirmed that pini/RG is in the range 3–8 for all outcomes resulting in satellite formation, as predicted in Equation (4).

Figure 1.

Figure 1. (a) Distribution of the initial semilatus rectum, pini, and the secondary-to-primary mass ratio of the system, γX,G, formed after giant impact simulations (Arakawa et al. 2019). The yellow region indicates the range of pini/RG (see relation (4)). (b) Initial spin periods of secondaries, PX,ini, formed after the giant impact simulations. The yellow region indicates the range of PX,ini from the giant impact simulations.

Standard image High-resolution image

Figure 1(b) shows the distribution of the initial spin period of secondaries after giant impacts, ${P}_{{\rm{X}},\mathrm{ini}}=2\pi /{\dot{\theta }}_{{\rm{X}},\mathrm{ini}}$, where ${\dot{\theta }}_{{\rm{X}},\mathrm{ini}}$ is the initial spin angular velocity of secondaries. Although it is difficult to analytically estimate PX,ini, the numerical results show that the initial spin period of the secondaries is in the range 3 hr ≲ PX,ini ≲ 18 hr. The lower limit of PX,ini is given by PX,ini > 2πcr, where

Equation (5)

is the critical spin angular velocity for rotational instability (e.g., Kokubo & Genda 2010); where ${ \mathcal G }$ is the gravitational constant, and ρ is the bulk density.

2.2. Initial Spins and Orbit of the Gonggong–Xiangliu System

In this section, we discuss the initial spins and orbit of the Gonggong–Xiangliu system from the perspective of the conservation of angular momentum. The total angular momentum of the system, Ltot, is

Equation (6)

and Ltot should be constant over time during the tidal evolution of the system. Here, MtotMG + MX is the total mass of the satellite system, and Ii and ${\dot{\theta }}_{i}$ are the moment of inertia and the spin angular velocity, respectively. Hereafter, the subscript i denotes the primary (i = G) or the secondary (i = X). For the case of undifferentiated bodies, the density distribution of the planetary interior is homogeneous, and the moment of inertia is approximately given by

Equation (7)

As the spin angular momentum of the secondary, Xiangliu, is negligibly small, we obtain the following equation:

Equation (8)

where ${\dot{\theta }}_{{\rm{G}},\mathrm{ini}}$ and ${\dot{\theta }}_{{\rm{G}},\mathrm{obs}}$ are the initial and observed (current) spin angular velocities of the primary, Gonggong; and pini and pobs are the initial and observed semilatus recta. Figure 2 shows the initial semilatus rectum, pini, as a function of the initial spin period of Gonggong, PG,ini, the observed spin period of Gonggong, PG,obs, and the secondary-to-primary mass ratio, γX,G.

Figure 2.

Figure 2. Initial semilatus rectum, pini, as a function of the initial spin period of Gonggong, PG,ini, the observed spin period of Gonggong, PG,obs, and the secondary-to-primary mass ratio, γX,G. The dashed lines indicate the initial location of the corotation radius, acorot. The yellow regions indicate the range of pini/RG inferred from the giant impact simulations (Figure 1(a)).

Standard image High-resolution image

We note that the current spin period of Gonggong has not yet been uniquely determined; the observation by Pál et al. (2016) suggested PG,obs = 22.4 hr or 44.8 hr, although the most likely light-curve solution is the double-peaked solution with a slight asymmetry. This corresponds to a spin period of PG,obs =44.8 hr. In contrast, Kiss et al. (2019) suggested that the most plausible solution for the system would be a single-peaked light curve in the visible range (i.e., PG,obs = 22.4 hr) caused by surface features rather than by a distorted shape, and the primary's equator and the secondary's orbit are coplanar.

We also calculated the initial location of the corotation radius, acorot, which is given by

Equation (9)

For an initial eccentricity of eini = 0, the tidal evolution of the satellite results in inward migration when the initial semimajor axis is aini < acorot. Then, the initial semilatus rectum is expected to be piniacorot to explain the current semimajor axis of the system, aobs, which is an order of magnitude larger than acorot.

Figure 2 indicates that the tidal evolution of the satellite results in outward migration when the secondary-to-primary mass ratio is approximately γX,G = 10−2 or larger. In contrast, the tidal evolution of the satellite would result in inward migration when γX,G ≪ 10−2 and pini/RG ≲ 4 (i.e., eini ≃ 0). Therefore, for the case of γX,G ≪ 10−2, the initial orbit of the Gonggong–Xiangliu system formed via a giant impact must have a finite eccentricity; otherwise, the satellite would migrate inward and become tidally disrupted. The minimum value of eini (and pini) is expected to depend on γX,G and PG,ini; we confirm this prediction in Sections 5 and 6.

2.3. Initial Conditions for Simulation of Tidal Evolution

We summarize the initial parameters for the calculations of tidal evolution in Table 2. In Section 5, we parameterize the radius of Xiangliu, RX, the initial eccentricity, eini, and the initial temperature, Tini. For simplicity, we set ρG = ρX, where ρG and ρX are the bulk densities of Gonggong and Xiangliu. We assume that the initial temperature of Xiangliu is equal to that of Gonggong (i.e., TG,ini = TX,ini = Tini) in Section 5. We also investigated the dependence of the outcomes of tidal evolution on the current spin period of Gonggong, PG,obs, the initial spin period of Xiangliu, PX,ini, the different settings of the initial temperature of Xiangliu (we set TX,ini = 120 K regardless of TG,ini), and the value of the reference viscosity (see Table 3), ηref, in Section 6.

Table 2. Initial Condition for Calculations of Tidal Evolution

SymbolParameterValueUnitComment
qini Initial periapsis distance2.1 × 103 km3 ≲ qini/RG ≲ 4 (Arakawa et al. 2019)
eini Initial eccentricity0.1, 0.2, ..., 0.80 < eini < 1 (Arakawa et al. 2019)
PG,ini Initial spin period of Gonggonghr ${P}_{{\rm{G}},\mathrm{ini}}=2\pi /{\dot{\theta }}_{{\rm{G}},\mathrm{ini}}$ is given by Equation (8)
PX,ini Initial spin period of Xiangliu 12 or Porb,ini hr3 hr < PX,ini < 18 hr (see Figure 1(b))
RG Radius of Gonggong600 kmKiss et al. (2019)
RX Radius of Xiangliu20, 40, ..., 120 km RX ≳ 18 km (see Kiss et al. 2019)
TG,ini Initial temperature of Gonggong120, 140, ..., 240K TG,ini is below the melting point of ice
TX,ini Initial temperature of Xiangliu T G,ini or 120KSee Section 6.3

Note. Parameters for standard runs are indicated by bold face.

Download table as:  ASCIITypeset image

Table 3. Material Properties Used in This Study

SymbolPropertyValueUnitReferences
c Specific heat capacity $8.8\left(T/{\rm{K}}\right)$ J kg−1 K−1 Hammond et al. (2016)
kth Thermal conductivity $0.48+488{\left(T/{\rm{K}}\right)}^{-1}$ W m−1 K−1 Hammond et al. (2016)
αexp Thermal expansion coefficient1 × 10−4 K−1 Kamata et al. (2019)
Ea Activation energy60kJ mol−1 Kamata et al. (2019)
Tref Reference temperature273KKamata et al. (2019)
ηref Reference viscosity at Tref 1014 or 1010 Pa sSee Section 6.4
μ Shear modulus3.33GPaRobuchon & Nimmo (2011)
α Andrade exponent0.33Rambaux et al. (2010)

Note. Parameters for standard runs are indicated by bold face.

Download table as:  ASCIITypeset image

In this study, we assume that the primary's equator and the secondary's orbit are coplanar, and the orbit of the satellite system is prograde (i.e., the spin–orbit angle is close to zero). This assumption is also consistent with the numerical simulations of Arakawa et al. (2019); however, their simulations do not consider the pre-impact spin of the target and impactor. Observations by Kiss et al. (2019) are consistent with both prograde and retrograde orbit. If the Gonggong–Xiangliu system is in retrograde orbit, the semimajor axis will decrease with increasing time. Thus the determination of the spin–orbit angle by future observations could provide critical constraints on the origin of the satellite system.

We also note that the observed coplanar orbit is for the present-day system, and this does not indicate that the system was born in a coplanar orbit. Although the giant impact simulations preferred a coplanar orbit, we should discuss the effect of nonzero initial inclination/obliquity on the tidal evolution in future studies. Several studies (e.g., Correia 2020; Renaud et al. 2021) pointed out that obliquity evolution can affect how bodies fall into and out of higher-order spin–orbit resonances.

3. Tidal Evolution Model

In this section, we briefly introduce the orbital evolution models (Section 3.1) and thermal evolution models (Section 3.2). We perform calculations of 4.5 Gyr tidal evolution with various initial conditions. We stopped numerical integration when the eccentricity reached e = 0.9 or the periapsis distance reached qorb = RG + RX.

3.1. Orbital Evolution

Tides are raised on both the primary and secondary. The tidal lag caused by friction leads to angular momentum exchange, which also leads to spin and orbital evolution. In this study, we use the tidal evolution equations following Boué & Efroimsky (2019). The orbit-averaged variations of the spin rates of the primary and secondary, ${\dot{\theta }}_{{\rm{G}}}$ and ${\dot{\theta }}_{{\rm{X}}}$, semimajor axis a, and eccentricity e are given by

Equation (10)

Equation (11)

Equation (12)

Equation (13)

Equation (14)

Equation (15)

Equation (16)

Equation (17)

where $n=\sqrt{{ \mathcal G }{M}_{\mathrm{tot}}/{a}^{3}}$ denotes the mean motion. The initial orbital period, Porb,ini, is ${P}_{{\rm{orb}},{\rm{ini}}}=2\pi /\sqrt{{ \mathcal G }{M}_{{\rm{tot}}}/{a}_{{\rm{ini}}}^{3}}$. Here, we assume that the orbital inclinations on both the primary's and secondary's equators are negligibly small. We note that Kiss et al. (2019) suggested that the orbit of the secondary would be coplanar with the equator of the primary. The coefficients ${{ \mathcal A }}_{{\rm{G}},q}$, ${{ \mathcal A }}_{{\rm{X}},q}$, ${{ \mathcal B }}_{{\rm{G}},q}$, and ${{ \mathcal B }}_{{\rm{X}},q}$, are given by

Equation (18)

Equation (19)

Equation (20)

Equation (21)

where G2,0,q and G2,1,q are the eccentricity functions (see Appendix A), and ${\tilde{k}}_{2,{\rm{G}}}$ and ${\tilde{k}}_{2,{\rm{X}}}$ are the Love numbers of Gonggong and Xiangliu, which depend on the tidal frequency (see Appendix B). We consider the terms G2,p,q (p = 0 or 1) when the following two conditions, ∣q∣ ≤ 200 and ${[{G}_{2,p,q}]}^{2}\geqslant {10}^{-20}$, are satisfied.

In our numerical integration, we calculate ${\dot{\theta }}_{{\rm{G}}}/n$ and ${\dot{\theta }}_{{\rm{X}}}/n$ instead of ${\dot{\theta }}_{{\rm{G}}}$ and ${\dot{\theta }}_{{\rm{X}}}$ (e.g., Cheng et al. 2014). The relation between ${dn}/{dt}$ and da/dt is

Equation (22)

and we obtain the following equations:

Equation (23)

Equation (24)

3.2. Thermal Evolution

In this study, we use the simplifying assumption of undifferentiated homogeneous bodies. We do not consider the internal temperature structure but calculate the effective temperature of the interior for simplicity 5 (e.g., Ojakangas & Stevenson 1986; Shoji et al. 2013). Then, the temperature evolution of the body i is given by the following equation:

Equation (25)

where ci is the specific heat capacity of the primary and secondary. The total heat generation rate within bodies, Qtot,i , is given by

Equation (26)

where Qtide,i , Qcon,i , and Qdec,i are the tidal heating, the conduction/convection cooling, and the decay heating terms, respectively. We note that the main heat source is not tidal heating but decay heating. We discuss the balance between conduction/convection cooling and decay heating for the primary in Appendix C.

3.2.1. Tidal Heating

The tidal heating rate, Qtide,i , is given by the sum of two terms:

Equation (27)

Equation (28)

Equation (29)

However, the contribution of the tidal heating is negligibly small when we consider the tidal evolution of 100 km sized satellites around 1000 km sized dwarf planets (e.g., Arakawa et al. 2019).

3.2.2. Conduction/Convection Cooling

The term for cooling by conduction/convection, Qcon,i , is given by

Equation (30)

Equation (31)

Equation (32)

where κi is the thermal diffusivity, which is given by

Equation (33)

and kth,i is the thermal conductivity. Reese et al. (1999) proposed that the Nusselt number, Nui , is given by the following scaling equation (see also Solomatov & Moresi 2000):

Equation (34)

Equation (35)

Equation (36)

where Θi and Rai are the Frank–Kamenetskii parameter and the Rayleigh number, respectively. The surface gravity, gi , is given by ${g}_{i}={ \mathcal G }{M}_{i}/{{R}_{i}}^{2}$, and Ea is the activation energy, Rgas is the gas constant, αexp is the thermal expansion coefficient, and ηi is the viscosity. The temperature dependence of the viscosity is given as follows (e.g., Goldsby & Kohlstedt 2001):

Equation (37)

where ηref is the reference viscosity and Tref is the reference temperature (Table 3). We set the surface temperature Tsurf = 40 K as assumed in previous studies on the thermal evolution of dwarf planets (e.g., Robuchon & Nimmo 2011; Kamata et al. 2019). The material parameters used in this study are summarized in Table 3.

We stress that the temperature evolution of bodies significantly impacts their spin/orbital evolution. As the viscosity is a function of the temperature, the complex Love number also depends on the temperature (see Appendix B). Then the complex Love number influences the spin/orbital evolution, which is calculated from Equations (10)–(17).

Although it is beyond the scope of this study, ices in undifferentiated icy bodies would not be pure water ice, and the partial melting of ice mixtures may have a great impact on the effective viscosity (e.g., Henning et al. 2009). For the case of ice mixtures containing 1% ammonia with respect to water, the effective viscosity is orders of magnitude lower than that of pure water ice when the temperature is above the ammonia–water eutectic temperature (176 K; e.g., Arakawa & Maeno 1994; Neveu & Rhoden 2017). We could mimic this effect by changing the reference viscosity, which is similar to the impact of crystal grain size on the reference viscosity (see Section 6.4).

3.2.3. Decay Heating

For 100–1000 km sized planetary bodies, the radioactive decay of long-lived isotopes is the principal heat source (e.g., Robuchon & Nimmo 2011). We assume that the elemental abundances of long-lived isotopes of the rocky parts of TNOs are equal to those of carbonaceous chondrites (Lodders 2003; Robuchon & Nimmo 2011). We consider four species as heat sources, namely 238U, 235U, 232Th, and 40K. The half-life, tHL,j , and the initial heat generation rate per unit mass of rock, H0,j , of the element j are shown in Table 4.

Table 4. Radioactive Species and Decay Data (Robuchon & Nimmo 2011)

Element j Half-life tHL,j (Myr) H0,j (W kg−1 of Rock)
238U44681.88 × 10−12
235U703.813.07 × 10−12
232Th14,0301.02 × 10−12
40K12772.15 × 10−11

Download table as:  ASCIITypeset image

Then the decay heating term is given by

Equation (38)

where frock is the mass fraction of rock in bodies. We set frock = 0.4 in this study. Kiss et al. (2019) determined that the bulk density of Gonggong is 1.75 × 103 kg m−3; our assumption of frock = 0.4 seems within an acceptable range if TNOs are made of a mixture of ices, rocks, and organic materials. In future studies, we will investigate the effect of the difference in frock on the thermal and orbital evolution.

4. Classification of Tidal Evolution Pathways

In this section, we present the statistics of the final state of the satellite systems. We classify the outcomes of tidal evolution into four types, namely 1:1 spin–orbit resonance (Type A), higher-order spin–orbit resonance (Type B), nonresonance (Type C), and collision with the primary (Type Z). No systems reached a dual-synchronous state, unlike the Pluto–Charon system. We introduce typical tidal evolution pathways of satellite systems in Appendix D.

4.1. Summary of the Final State

Figure 3 shows the summary of the final state for standard runs of our simulation. In our standard runs, we set PG,obs =22.4 hr, PX,ini = 12 hr, TX,ini = TG,ini = Tini, and ηref =1014 Pa s. We change the radius of Xiangliu, RX, the initial eccentricity, eini, and the initial temperature, Tini. The different markers represent the different final states of the satellite system classified according to the spin state of Xiangliu, semimajor axis, and eccentricity.

Figure 3.

Figure 3. Summary of the final state of the standard runs of our simulation (i.e., PG,obs = 22.4 hr, PX,ini = 12 hr, TX,ini = TG,ini = Tini, and ηref = 1014 Pa s). We changed the radius of Xiangliu, RX, the initial eccentricity, eini, and the initial temperature, Tini, as parameters. The different markers represent the different final states of the tidal evolution classified according to the spin state of Xiangliu, semimajor axis, and eccentricity.

Standard image High-resolution image

Figure 3(a) shows the outcomes of the tidal evolution for the case of RX = 20 km. In this setting, Xiangliu migrates inward and finally collides with Gonggong when the initial eccentricity is eini ≤ 0.2 (Type Z). The critical value of eini required for the outward migration of Xiangliu is clearly dependent on its radius; smaller satellites have a larger critical value of eini. This trend is consistent with our analytic arguments in Section 2.2.

.Figure 3(b) shows the outcomes of the tidal evolution for the case of RX = 40 km. In this setting, some runs with large eini and high Tini result in extremely eccentric systems with e > 0.9 within 4.5 Gyr (Types Bx/Cx). The parameter space resulting in Types Bx/Cx is wider for larger values of RX. In addition, the runs with eini ≳ 0.7 tend to become extremely eccentric when RX ≥ 40 km.

Figure 3(f) shows the outcomes of the tidal evolution for the case of RX = 120 km. With this setting, runs with small eini and high Tini result in Type A, i.e., Xiangliu is trapped in 1:1 spin–orbit resonance at t = 4.5 Gyr. In general, runs with small eini and high Tini tend to turn into Type A after tidal evolution unless their initial eccentricity is too small to satisfy the condition for the outward migration of Xiangliu (see Figures 3(c) and (d)).

We did not find any clear trends for the boundary between Types B and C. In other words, it is difficult to predict whether the secondary is trapped in a higher-order spin–orbit resonance or not trapped in resonances. The fractions of Type B and Type C were 65/336 = 19% and 52/336 = 15%, respectively, according to all our runs with the standard setting (see Table 5). We note that we did not consider the realistic probability distributions of RX, eini, and Tini for satellite systems around TNOs.

Table 5. Statistics of the Final State for Standard Runs of Our Simulation (see also Figure 3)

 (a) RX = 20 km(b) 40 km(c) 60 km(d) 80 km(e) 100 km(f) 120 kmTotal
Type A0 (0%)3 (5%)12 (21%)18 (32%)27 (48%)31 (55%)91 (27%)
Type B26 (46%)12 (21%)9 (16%)7 (13%)9 (16%)2 (4%)65 (19%)
Type C16 (29%)18 (32%)12 (21%)5 (9%)0 (0%)1 (2%)52 (15%)
Type Bx0 (0%)4 (7%)6 (11%)14 (25%)17 (30%)16 (29%)57 (17%)
Type Cx0 (0%)5 (9%)10 (18%)5 (9%)3 (5%)6 (11%)29 (9%)
Type Z14 (25%)14 (25%)7 (13%)7 (13%)0 (0%)0 (0%)42 (13%)

Download table as:  ASCIITypeset image

Table 5 shows the statistics of the final state for standard runs of our simulation. Figure 3 shows that the fraction of each type is clearly dependent on the radius of the secondary, RX. The fractions of Type A, Type Bx, and Type Cx increase with increasing RX. In contrast, the fractions of Type B, Type C, and Type Z decrease with increasing RX.

Our results also indicate that other satellites around 1000 km sized TNOs may not be in 1:1 spin–orbit resonance. In the Haumea–Hi'iaka system, the spin period of the secondary, Hi'iaka, is approximately 120 times shorter than its orbital period (Hastings et al. 2016). Although the fast spin of Hi'iaka could be explained by a scenario in which the current satellite system was formed via the catastrophic disruption of the first-generation moon and subsequent reaccretion of fragments (e.g., Schlichting & Sari 2009; Ćuk et al. 2013), we could also explain the fast spin of Hi'iaka without the disruption event. We note, however, that the radius of Hi'iaka is 150 km (Ragozzine & Brown 2009), and the fraction of the nonresonant case (Type C) is small in our standard simulations for the Gonggong–Xiangliu system with RX ≳ 100 km. Moreover, the spin period of the primary, Haumea, is PG,obs = 3.9 hr (Rabinowitz et al. 2006), which is an order of magnitude shorter than that of Gonggong. In addition, the nonspherical and highly elongated shape of Haumea (e.g., Ortiz et al. 2017) may also play an important role in their tidal evolution. We will apply our tidal evolution model to the Haumea–Hi'iaka system to unveil its spin–orbit evolution in future studies.

4.2. Condition for Spin–Orbit Coupling

Section 4.1 shows some cases in which the secondary is not in spin–orbit resonance after a 4.5 Gyr orbital evolution (i.e., Type C). Here we discuss the condition for capture into spin–orbit resonance.

Appendix B shows that the imaginary part of the Love number, $\mathrm{Im}[{\tilde{k}}_{2,{\rm{X}}}({\omega }_{{\rm{X}},q})]$, is a minimum/maximum when ${\omega }_{{\rm{X}},q}\simeq \pm {\left({\mu }_{\mathrm{eff},{\rm{X}}}{\tau }_{{\rm{A}},{\rm{X}}}\right)}^{-1}$, where ${\omega }_{{\rm{X}},q}=(2+q)n-2{\dot{\theta }}_{{\rm{X}}}$ is the tidal frequency. When the secondary is in the spin–orbit resonance of order q, the tidal frequency satisfies $-{\left({\mu }_{\mathrm{eff},{\rm{X}}}{\tau }_{{\rm{A}},{\rm{X}}}\right)}^{-1}\lt {\omega }_{{\rm{X}},q}\lt {\left({\mu }_{\mathrm{eff},{\rm{X}}}{\tau }_{{\rm{A}},{\rm{X}}}\right)}^{-1}$, and the spin angular velocity of the secondary, ${\dot{\theta }}_{{\rm{X}}}$, is given by

Equation (39)

where ∣x∣ < 1 is a dimensionless parameter. For ∣x∣ ≪ 1, the imaginary part of the Love number is given by

Equation (40)

Here, we assume that the following relations are satisfied for integers $q^{\prime} \ne q$ (see Section 3.1 for the definitions of ${{ \mathcal A }}_{{\rm{X}},q}$ and ${{ \mathcal B }}_{{\rm{X}},q}$):

Equation (41)

Under this assumption, $d{\dot{\theta }}_{{\rm{X}}}/{dt}$, (da/dt)X, and $d({\dot{\theta }}_{{\rm{X}}}/n)/{dt}$ near the spin–orbit resonance are approximately given by

Equation (42)

Equation (43)

Equation (44)

Assuming that (da/dt)G > 0, the required conditions for a stable equilibrium around ${\dot{\theta }}_{{\rm{X}}}\simeq (2+q)n/2\,+\,x/({\mu }_{\mathrm{eff},{\rm{X}}}{\tau }_{{\rm{A}},{\rm{X}}})$ with ∣x∣ ≪ 1 are

Equation (45)

Equation (46)

On the other hand, for the case of ${\left({da}/{dt}\right)}_{{\rm{G}}}\lt 0$, the required conditions for a stable equilibrium are

Equation (47)

Equation (48)

Except for the case of q = 0, ${\left[{G}_{\mathrm{2,0},q}(e)\right]}^{2}\to 0$ when e → 0 (see Appendix A). Thus, a finite eccentricity is needed for capture into higher-order spin–orbit resonances with q ≠ 0.

5. Distributions of Final Spin/Orbital Properties of the System

In this section, we present the distributions of the final spin/orbital properties of the system.

5.1. Semimajor Axis and Eccentricity

Figure 4 shows the distributions of the final eccentricity and semimajor axis, efin and afin. The observed values for the Gonggong–Xiangliu system are efin = 0.3 and afin/RG =24,000 km/600 km = 40; the green stars indicate the observed values.

Figure 4.

Figure 4. Distributions of the final eccentricity and semimajor axis, efin and afin, for standard runs of our simulation. The different markers represent the different final states of the tidal evolution classified according to the spin state of Xiangliu and the final semimajor axis and eccentricity. For the cases of Type Bx/Cx/Z, we stopped numerical integrations when the eccentricity reached e = 0.9 or when the periapsis distance reached qorb = RG + RX.

Standard image High-resolution image

We found that the final eccentricity is efin ≲ 0.05 for Type A. Xiangliu has a moderate eccentricity of eobs = 0.3 (Kiss et al. 2019), and therefore it would not be in 1:1 spin–orbit resonance. The final eccentricity of satellite systems classified as Type B and Type C is in the wide range 0 < efin < 0.9. We note, however, that the fraction of systems whose efin is comparable to eobs depends strongly on the radius of Xiangliu. The fraction of systems whose final eccentricity is in the range 0.2 ≤ efin ≤ 0.5 is as follows: 10/56 = 18% for the case of RX = 20 km, 5/56 = 8.9% for 40 km, 4/56 = 7.1% for 60 km, 3/56 = 5.4% for 80 km, 4/56 = 7.1% for 100 km, and 0/56 = 0% for 120 km. Thus, the radius of Xiangliu is likely not larger than 100 km.

We also found a clear relation between afin and efin. The dashed lines in Figure 4 are given by

Equation (49)

for each Rs. We explain the reason why afin and efin are given by Equation (49) in Section 5.2. Figure 4 shows that the radius of Xiangliu might be close to 100 km from the point of view of pfin. We note that we should consider the dependence of pfin on many properties of the system, such as the viscoelastic properties of icy bodies and their thermal histories (see Sections 6.3 and 6.4).

Our results suggest that not only Gonggong–Xiangliu but also the Quaoar–Weywot system would not be in 1:1 spin–orbit resonance. Fraser et al. (2013) reported that the eccentricity of the Quaoar–Weywot system is e = 0.13–0.16. This eccentricity indicates that the system might be classified as Type B or Type C, because systems classified as Type A have a small eccentricity of e ≲ 0.05 in general.

5.2. Analytic Solution of the Semilatus Rectum

Figure 4 shows that the final semilatus rectum, pfin, hardly depends on the final eccentricity, efin. In this section, we derive an analytic solution for the evolution of the semilatus rectum.

The time derivative of porb is given by the following equation:

Equation (50)

Here, we assume that the contribution of the tidal torques caused by the secondary is negligible:

Equation (51)

Assuming that the spin period of the primary is sufficiently shorter than the orbital period (i.e., ${\dot{\theta }}_{{\rm{G}}}\gg n$), we can rewrite the term ${\sum }_{q}{{ \mathcal A }}_{{\rm{G}},q}$ as follows:

Equation (52)

We found that ${\sum }_{q}{\left[{G}_{\mathrm{2,0},q}(e)\right]}^{2}$ is approximately given by

Equation (53)

in the range 0 ≤ e ≤ 0.8 (see Figure A2). Therefore, the time evolution of porb is approximately given by the following equation:

Equation (54)

and we obtain porb = porb(t) as

Equation (55)

which does not include the eccentricity, e, in the equation.

Figure 5 shows the final semilatus rectum, pfin, as a function of the radius of the secondary, RX. We only used the data classified as Type A/B/C in the figure. We plotted the average value and twice the standard error of pfin for each RX. Equation (55) shows that the final semilatus rectum is proportional to ${R}_{{\rm{X}}}^{6/13};$ our numerical results are consistent with the theoretical prediction.

Figure 5.

Figure 5. Final semilatus rectum, pfin, as a function of the radius of the secondary, RX. The dashed line represents the least-squares fit, ${p}_{\mathrm{fin}}/{R}_{{\rm{G}}}=3.9{\left({R}_{{\rm{X}}}/\mathrm{km}\right)}^{6/13};$ the error bars are twice the standard deviation for each RX.

Standard image High-resolution image

5.3. Spin Period of the Secondary

Figure 6 shows the distribution of the final eccentricity and spin period of the secondary, efin and PX,fin, for standard runs of our simulation. The spin period of Xiangliu at t = 4.5 Gyr is in the range 10 hr ≲ PX,fin ≲ 103 hr, and the range of PX,fin depends on RX. In the case of RX = 20 km, the final spin period of Xiangliu is in the range 10 hr ≲ PX,fin ≲ 102 hr and is always shorter than the orbital period, Porb,fin. In contrast, for RX = 100 and 120 km, the final spin period of Xiangliu is in the range 102 hr ≲ PX,fin ≲ 103 hr, and PX,fin is significantly longer than the initial spin period, PX,ini = 12 hr. Thus, the determination of the spin period of Xiangliu by future observations and light-curve analyses is the key to estimating its radius.

Figure 6.

Figure 6. Distribution of the final eccentricity and spin period of the secondary, efin and PX,fin, for the standard runs of our simulation. The different markers represent the different final states of the tidal evolution classified according to the spin state of Xiangliu and the final semimajor axis and eccentricity. For the cases of Types Bx/Cx/Z, we stopped numerical integrations when the eccentricity reached e = 0.9 or when the periapsis distance reached qorb = RG + RX.

Standard image High-resolution image

The final spin period of Xiangliu is close to the initial spin period, PX,finPX,ini, for the case of Type C. The spin period of satellites formed via a giant impact should provide abundant information regarding the condition of the colliding TNOs, including the mass ratio, the spin state before the impact, and the (un)differentiated state. Therefore, we want to determine the spin period of satellites from astronomical observations for Gonggong–Xiangliu and for other satellite systems around large TNOs.

5.4. Final Eccentricity and the Initial Condition

Figure 7 shows the final eccentricity of the system, efin, for standard runs of our simulation. We note that we do not plot the results classified into Type Z.

Figure 7.

Figure 7. Color maps for the final eccentricity of the system, efin, for standard runs of our simulation.

Standard image High-resolution image

Figure 7 shows that satellite systems with an initial eccentricity of 0.2 ≤ eini ≤ 0.5 tend to a final eccentricity of 0.2 ≤ efin ≤ 0.5. When the initial eccentricity is eini ≤ 0.1, the final eccentricity at t = 4.5 Gyr is typically efin ≤ 0.1 or the system becomes Type Z. When the initial eccentricity is eini ≥ 0.6, on the other hand, the final eccentricity at t = 4.5 Gyr is typically efin ≥ 0.5 for the case of RX ≤ 60 km, and efin ≥ 0.9 or efin ≤ 0.1 for the case of RX ≥ 80 km.

Therefore, we conclude that the initial eccentricity of the Gonggong–Xiangliu system after the moon-forming giant impact may be in the range 0.2 ≲ eini ≲ 0.5. This nonzero eini is also consistent with the fact that eini ≳ 0.2 is needed to avoid the inward migration of Xiangliu (see Figure 3).

Debris disks can be formed by a giant impact (Canup 2005), collisional disruption of a pre-existing satellite (Hyodo & Charnoz 2017), or tidal disruption of a passing body (Hyodo et al. 2017). If a single large satellite is formed via accretion of debris disk materials around Gonggong, the initial eccentricity becomes small: e ≲ 0.1 (e.g., Ida et al. 1997; Kokubo et al. 2000). In this case, we cannot reproduce the observed eccentricity of the system. Thus, the Gonggong–Xiangliu system was more likely to be born as an "intact moon" rather than "disk-origin moon," and their formation process is similar to that of the Pluto–Charon system (e.g., Canup 2005; Sekine et al. 2017; Arakawa et al. 2019).

6. Discussion

In Section 5, we show the distributions of the final spin/orbital properties of the Gonggong–Xiangliu system for our standard setting (i.e., PG,obs = 22.4 hr, PX,ini = 12 hr, TX,ini =TG,ini = Tini, and ηref = 1014 Pa s). In this section, we discuss the effect of a nonstandard setting on the results of the tidal evolution. We investigate four cases in this section: (a) a slow spin of Gonggong (PG,obs = 44.8 hr instead of 22.4 hr), (b) an initially tidally synchronized Xiangliu (PX,ini = Porb,ini instead of 12 hr), (c) a cold start for Xiangliu (TX,ini = 120 K instead of TG,ini), and (d) undifferentiated bodies with low reference viscosity (ηref = 1010 Pa s instead of 1014 Pa s). Figures 811 are the results of the nonstandard runs of our simulations. In these simulations, we set RX = 60 km.

Figure 8.

Figure 8. Summary of the final state for nonstandard runs of our simulation. We changed the initial eccentricity, eini, and the initial temperature (of Gonggong), Tini (TG,ini), as parameters. The different markers represent the different final states of the tidal evolution classified according to the spin state of Xiangliu and the final semimajor axis and eccentricity.

Standard image High-resolution image
Figure 9.

Figure 9. Distribution of the final eccentricity and semimajor axis, efin and afin, for nonstandard runs of our simulation.

Standard image High-resolution image
Figure 10.

Figure 10. Distribution of the final eccentricity and spin period of the secondary, efin and PX,fin, for nonstandard runs of our simulation.

Standard image High-resolution image
Figure 11.

Figure 11. Color maps for the final eccentricity of the system, efin, for nonstandard runs of our simulation.

Standard image High-resolution image

6.1. Slow-spinning Gonggong

In our standard models, we set the current spin period of Gonggong as PG,obs = 22.4 hr. The spin period of Gonggong was obtained from the light-curve analysis by Pál et al. (2016) and has two possible values: PG,obs = 22.4 hr or 44.8 hr. In Section 2.2, we discuss the impact of PG,obs on the location of the initial corotation radius. We showed that a large eini is necessary to avoid the inward migration of Xiangliu when PG,obs = 44.8 hr; however, a nonzero value of eini is required even for the case of PG,obs = 22.4 hr. As we cannot exclude the possibility of PG,obs = 44.8 hr from the light-curve analyses, we discuss the case in this section.

Panels (a) in Figures 811 show the results for the case of PG,obs = 44.8 hr. Figure 8(a) shows that the satellite system results in Type Z for an initial eccentricity of eini ≤ 0.4. The critical value of eini to avoid the inward migration of Xiangliu is significantly larger than that for our standard runs (see Figure 3(c)), which is consistent with our prediction from the analytical estimate of the corotation radius in Section 2.2.

Figure 9(a) shows the distribution of efin and afin. The final eccentricity of satellite systems classified as Type B and Type C is in the wide range 0 < efin < 0.9, as in the case shown in Figure 4(c). The final semilatus rectum was slightly lower than the observed value for the Gonggong–Xiangliu system. Figure 10(a) shows the distribution of efin and PX,fin. The general trend is similar to that of the standard shown in Figure 6(c).

Figure 11(a) is the color map for efin. In the case of PG,obs = 44.8 hr, satellite systems with an initial eccentricity of eini ≃ 0.6 tend to a final eccentricity of 0.2 ≤ efin ≤ 0.5. Systems with large eini tend to result in a large efin; this trend is similar to that observed in the standard runs (see Figure 7(c)); although the suitable range of eini to reproduce the observed value for the Gonggong–Xiangliu system is different.

6.2. Initially Tidally Synchronized Xiangliu

In our standard models, we set the initial spin period of Xiangliu as PX,ini = 12 hr. This assumption is based on numerical simulations of moon-forming giant impacts of 1000 km sized TNOs (see Figure 1(b)). In this case, PX,iniPorb,ini and a large number of runs are trapped in higher-order spin–orbit resonances (see Figure 3). In this section, we instead set PX,ini = Porb,ini. Although the assumption of PX,ini = Porb,ini seems unrealistic, we discuss the effects of the initially synchronized Xiangliu on the tidal evolution of the system to better understand our results.

Panels (b) in Figures 811 show the results for the case of PX,ini = Porb,ini. Figure 8(b) shows that the final state of the system is not Type A (i.e., 1:1 spin–orbit resonance) in some cases even if the initial spin period of the secondary is PX,ini = Porb,ini.

Figure 9(b) shows the distribution of efin and afin. For the case of PX,ini = Porb,ini, the final eccentricity is in the wide range, even for Type A (0 < efin < 0.7). We note, however, that the assumption of PX,ini = Porb,ini is unrealistic if they are formed via giant impacts (see Figure 1(b)). Figure 10(b) shows the distribution of efin and PX,fin. In this case, the final spin period of Xiangliu lies in the range 102 hr ≲ PX,fin ≲ 103 hr. In addition, all satellite systems with efin ≤ 0.6 are classified as Type A.

Figure 11(b) is the color map for efin. In the case of PX,ini = Porb,ini, satellite systems with an initial eccentricity of 0.2 ≤ eini ≤ 0.4 tend to a final eccentricity of 0.2 ≤ efin ≤ 0.5. This result is similar to that of our standard runs (see Figure 7).

6.3. Cold Start for Xiangliu

In our standard models, we assumed that the initial temperatures of Gonggong and Xiangliu are the same: TG,ini = TX,ini = Tini. Although this assumption is somewhat natural, we can imagine a case in which the initial temperature of Xiangliu is lower than that of Gonggong. The motivation for this setting is as follows.

Giant impact simulations of 1000 km sized TNOs (e.g., Canup 2005; Sekine et al. 2017; Arakawa et al. 2019) revealed that satellites are directly formed as intact fragments of the impactor ("intact moons"). As the mass of the impactor is several times smaller than that of the target body, the cooling rate of the impactor would be faster than that of the target, and the temperature of the impactor might be lower than that of the target upon collision. Although we need to assess the effect of the impact heating on the initial temperatures of the primary and secondary, the setting in which the initial temperature of Xiangliu is lower than that of Gonggong seems to be reasonable.

Panels (c) in Figures 811 show the results for the case of TX,ini = 120 K. Figure 8(c) shows that no systems classified as Type A are formed in this case.

Figure 9(c) shows the distribution of efin and afin, and Figure 10(c) shows the distribution of efin and PX,fin. These distributions are generally similar to those for our standard runs (Figures 4(c) and 6(c)); however, no systems result in 1:1 spin–orbit resonance.

Figure 11(c) is the color map for efin. In the case of TX,ini = 120 K, satellite systems with an initial eccentricity of 0.2 ≤ eini ≤ 0.4 tend to a final eccentricity of 0.2 ≤ efin ≤ 0.5. This result is also similar to those of our standard runs (see Figure 7(c)).

6.4. Undifferentiated Bodies Made of Soft Ice

In our standard models, we set the reference viscosity of icy bodies as ηref = 1014 Pa s. This value is widely used in studies of differentiated icy bodies, including of the thermal evolution of Pluto (e.g., Kamata et al. 2019), the orbital evolution of Pluto–Charon (e.g., Renaud et al. 2021), and the tidal heating of Europa and Titan (e.g., Tobie et al. 2005).

However, we note that the reference viscosity of icy bodies depends strongly on the grain size of ice crystals (e.g., Goldsby & Kohlstedt 2001; Kubo et al. 2006), and ηref for ice mixed with a small volume fraction of dust grains may be orders of magnitude lower than the canonical value for differentiated ice bodies (e.g., Kubo et al. 2009). When the ice viscosity is controlled by diffusion creep, the reference viscosity, ηref, at the reference temperature, Tref = 273 K, is given by (Kalousová et al. 2016)

Equation (56)

where dice is the grain size of the ice crystals, A = 3.3 ×10−10 Pa−1 m2 s−1 is the pre-exponential constant, Rgas =8.31 J K−1 mol−1 is the gas constant, and Ea is the activation energy. Assuming Ea = 60 kJ mol−1 (Kamata et al. 2019), we obtain the following relation between dice and ηref:

Equation (57)

When undifferentiated icy bodies are mixtures of micron-sized dust grains and matrix ices, the grain size of ice crystals would be maintained in the micron size owing to the Zener pinning effect (Kubo et al. 2009). Thus, the reference viscosity of ηref ∼ 1010 Pa s may be somewhat reasonable for undifferentiated icy bodies.

Panels (d) in Figures 811 show the results for the case of ηref = 1010 Pa s. Figure 8(d) shows that runs with a small eini and high Tini tend to result in Type A.

Figure 9(d) shows the distribution of efin and afin, and Figure 10(d) shows the distribution of efin and PX,fin. These distributions are similar to that of our standard runs. The final semilatus rectum, pfin, for the case of ηref = 1010 Pa s is not significantly different from that for the standard case of ηref = 1014 Pa s. This is because the temperature of Gonggong, TG, also depends on ηref; and the viscosity of Gonggong, ηG = η(TG), is not strongly dependent on ηref between the two settings, ηref = 1010 and 1014 Pa s. Figure D8 shows the thermal and orbital evolutions of Gonggong and Xiangliu for the case of ηref = 1010 Pa s, Tini = 140 K, and eini = 0.4.

Figure 11(d) is the color map for efin. In the case of ηref = 1010 Pa s, satellite systems with an initial temperature of Tini ≥ 160 K tend to result in Type A or Type Bx/Cx (i.e., a final eccentricity of efin ≤ 0.1 or efin ≥ 0.9). Therefore, the initial temperature of Tini ≤ 140 K might be suitable for explaining the formation of moderately eccentric satellite systems when ηref = 1010 Pa s.

7. Summary

Recent astronomical observations by Kiss et al. (2019) revealed that (225088) Gonggong, a 1000 km sized trans-Neptunian dwarf planet, hosts an eccentric satellite, Xiangliu, whose eccentricity is approximately eobs = 0.3. As the majority of known satellite systems around large TNOs have circular orbits with eobs ≲ 0.1, the observed eccentricity of the Gonggong–Xiangliu system may reflect the singular properties of this system. However, no detailed analysis of the orbital evolution of this system has been conducted so far.

In this study, we investigated the secular tidal evolution of the Gonggong–Xiangliu system under the simplifying assumption of homogeneous bodies, assuming that their initial orbital properties are those obtained from a giant impact (see Table 2). We conducted the simulations of coupled thermal–orbital evolution using the Andrade viscoelastic model and by including higher-order eccentricity functions up to ∣q∣ ≤ 200 for ${\left[{G}_{\mathrm{2,0},q}(e)\right]}^{2}$ and ${\left[{G}_{\mathrm{2,1},q}(e)\right]}^{2}$. Our findings are summarized as follows.

  • 1.  
    A re-evaluation of giant impact simulations by Arakawa et al. (2019) revealed that the initial semilatus rectum, ${p}_{{\rm{ini}}}={a}_{{\rm{ini}}}\left(1-{e}_{{\rm{ini}}}^{2}\right)$, lies in the range 3 < pini/RG < 8 when a giant impact forms a satellite as an intact fragment. In addition, the initial spin period of secondaries after giant impacts, PX,ini, lies in the range 3 hr ≲PX,ini ≲ 18 hr (see Figure 1).
  • 2.  
    In Section 4, we reported four typical tidal evolution pathways of satellite systems: 1:1 spin–orbit resonance (Type A), higher-order spin–orbit resonance (Type B), nonresonance (Type C), and collision with the primary (Type Z).
  • 3.  
    Figure 3 shows the outcomes of the tidal evolution. We found that Xiangliu migrates inward and finally collides with Gonggong (i.e., Type Z) when the initial eccentricity is too small. The critical value of eini required for the outward migration of Xiangliu is clearly dependent on the radius of Xiangliu, RX; smaller satellites have a larger critical value of eini. In addition, runs with small eini and high Tini tend to turn into Type A as a consequence of tidal evolution. Table 5 shows the statistics of the final state for the standard runs of our simulation. The fraction of each type is clearly dependent on the radius of the secondary.
  • 4.  
    We found that the final eccentricity is efin ≲ 0.05 for Type A. Therefore Xiangliu would not be in 1:1 spin–orbit resonance (i.e., Type A) because it has a moderate eccentricity of eobs = 0.3 (Kiss et al. 2019). The final eccentricity of satellite systems classified as Type B and Type C is in the wide range 0 < efin < 0.9. However, the fraction of systems whose final eccentricity is in the range 0.2 ≤ efin ≤ 0.5 depends strongly on the radius of Xiangliu, and no systems with 0.2 ≤ efin ≤ 0.5 were formed for the case of 120 km. These results suggest that the radius of Xiangliu would not be larger than 100 km.
  • 5.  
    We also found a significant relationship between afin and efin. In Section 5.2, we derived a simple formula for pfin; the final semilatus rectum is given by Equation (55). Figures 4 and 5 show that the radius of Xiangliu might be close to 100 km from the point of view of pfin.
  • 6.  
    From the aspect of the evolution of the eccentricity and semilatus rectum, our findings suggest that the radius of Xiangliu is approximately 100 km. This value is consistent with the estimate of Kiss et al. (2017) and indicates that Gonggong and Xiangliu have similar albedos.
  • 7.  
    Figure 6 shows the distributions of the final eccentricity and spin period of the secondary, efin and PX,fin, for standard runs of our simulation. The spin period of Xiangliu at t = 4.5 Gyr is in the range 10 hr ≲ PX,fin ≲ 103 hr; the range of PX,fin may depend on RX. In the case of RX = 20 km, the final spin period of Xiangliu is in the range 10 hr ≲ PX,fin ≲ 102 hr and is always shorter than the orbital period, Porb,fin. In contrast, for RX = 100 km and 120 km, the final spin period of Xiangliu is in the range 102 hr ≲ PX,fin ≲ 103 hr, which is significantly longer than the initial spin period, PX,ini = 12 hr. Thus, the determination of the spin period of Xiangliu by future observations and by light-curve analyses is necessary to evaluate the radius of Xiangliu.
  • 8.  
    We found that satellite systems with an initial eccentricity of 0.2 ≤ eini ≤ 0.5 tend to achieve a final eccentricity of 0.2 ≤ efin ≤ 0.5 (Figure 7). In contrast, for initial eccentricity of eini ≤ 0.1, the final eccentricity at t = 4.5 Gyr is typically efin ≤ 0.1, or the system become Type Z. For initial eccentricity of eini ≥ 0.6, on the other hand, the final eccentricity at t = 4.5 Gyr is typically efin ≥ 0.5 for the case of RX ≤ 60 km, and efin ≥ 0.9 or efin ≤ 0.1 for the case of RX ≥ 80 km. Therefore, the initial eccentricity of the Gonggong–Xiangliu system after the moon-forming giant impact may have been in the range 0.2 ≲ eini ≲ 0.5.

Our results highlight the importance of simulations of coupled thermal–orbital evolution using a realistic viscoelastic model with higher-order eccentricity functions. We also note that both the thermal history and the strength of tides depend strongly on the internal structure of the bodies. We assumed that the primary and secondary are undifferentiated homogeneous bodies for simplicity; however, 1000 km sized TNOs, including Gonggong, might actually be differentiated. In future studies, we should calculate the coupled thermal–orbital evolution of differentiated bodies. In this case, the effects of subsurface oceans are also of great interest. The differentiated state of large TNOs is closely associated with their accretion history, which is the key to understanding planet formation in the outer solar system. Thus, more detailed analyses of the coupled thermal–orbital evolution of satellite systems around large TNOs are necessary.

We thank two anonymous reviewers for their thoughtful comments. S.A. was supported by JSPS KAKENHI grants Nos. JP17J06861 and JP20J00598. R.H. was supported by JSPS KAKENHI grants Nos. JP17J01269 and JP18K13600. R.H. also acknowledges JAXA's International Top Young program. H.G. acknowledges the financial support of MEXT KAKENHI grant No. JP17H06457. This work was supported by the Publications Committee of NAOJ.

Appendix A: Eccentricity Function

The eccentricity functions are the Cayley expansions for the solutions of the Keplerian motion (Cayley 1861). The eccentricity functions, G2,p,q (e), can be calculated as follows (see Ferraz-Mello 2013, for details):

Equation (A1)

where r is the distance between the primary and secondary, ν is the true anomaly, and is the mean anomaly. The relation between r and ν is

Equation (A2)

and the mean anomaly is given by

Equation (A3)

Ferraz-Mello (2013) noted that the Fourier series solutions of G2,p,q (e) (e.g., Efroimsky 2012a; Renaud et al. 2021) are based on the Taylor expansion of the solution of Kepler's equation, whose convergence radius is e* = 0.6627434 (see Wintner 1941; Hagihara 1970; Murray & Dermott 1999). Therefore we should calculate G2,p,q (e) from direct integration of Equation (A1) when we consider the tidal evolution of highly eccentric systems with e ≫ 0.5. Figure A1 shows the squares of the eccentricity functions, ${\left[{G}_{\mathrm{2,0},q}(e)\right]}^{2}$ and ${\left[{G}_{\mathrm{2,1},q}(e)\right]}^{2}$, as functions of the eccentricity e and the order q.

Figure A1.

Figure A1. Squares of the eccentricity functions, ${\left[{G}_{\mathrm{2,0},q}(e)\right]}^{2}$ and ${\left[{G}_{\mathrm{2,1},q}(e)\right]}^{2}$ for different eccentricities: (a) e = 0.2, (b) e = 0.4, (c) e = 0.6, and (d) e = 0.8.

Standard image High-resolution image

Table 2 of Renaud et al. (2021) shows the series solutions of ${\left[{G}_{2,p,q}(e)\right]}^{2}$, with eccentricity terms up to and including e10. In the range ∣q∣ ≤ 2, the series solutions of ${\left[{G}_{\mathrm{2,0},q}(e)\right]}^{2}$ and ${\left[{G}_{\mathrm{2,1},q}(e)\right]}^{2}$ are given by Boué & Efroimsky (2019):

Equation (A4)

Equation (A5)

Equation (A6)

Equation (A7)

Equation (A8)

Equation (A9)

Equation (A10)

Equation (A11)

Figure A2 shows the summations of ${\left[{G}_{\mathrm{2,0},q}(e)\right]}^{2}$ and ${\left[{G}_{\mathrm{2,1},q}(e)\right]}^{2}$ from q = −200 to +200. This wide range of q allows sufficient summation convergence. Both ${\sum }_{q}{\left[{G}_{\mathrm{2,0},q}(e)\right]}^{2}$ and ${\sum }_{q}{\left[{G}_{\mathrm{2,1},q}(e)\right]}^{2}$ are approximately given by

Equation (A12)

for the range 0 ≤ e ≤ 0.8.

Figure A2.

Figure A2. Summations of ${\left[{G}_{\mathrm{2,0},q}(e)\right]}^{2}$ and ${\left[{G}_{\mathrm{2,1},q}(e)\right]}^{2}$ from q = −200 to +200. The dashed line indicates ${\left(1-{e}^{2}\right)}^{-6}$, and the summations of ${\left[{G}_{\mathrm{2,0},q}(e)\right]}^{2}$ and ${\left[{G}_{\mathrm{2,1},q}(e)\right]}^{2}$ are well approximated by ${\left(1-{e}^{2}\right)}^{-6}$ in the range 0 ≤ e ≤ 0.8.

Standard image High-resolution image

Appendix B: Rheological Model

B.1. Complex Shear Modulus

Viscoelastic models are widely used to describe the tidal response of small icy bodies. The Andrade model is an empirical model based on experimental results for the viscous flow of metals and ice (e.g., Andrade 1910; Glen 1955). The model is widely used to compute tidal dissipation within icy satellites. For example, Shoji et al. (2013) calculated the tidal heating of Enceladus using the Andrade model. Neveu & Rhoden (2019) also performed numerical simulations of coupling thermal, geophysical, and orbital evolution using the model.

The Andrade viscoelastic model contains an anelastic part, and the effect of elasticity decreases in the high-frequency region. The complex shear modulus, $\tilde{\mu }(\omega )$, is expressed in terms of a creep function as follows:

Equation (B1)

where $\tilde{J}(\omega )$ is the creep function and ω is the angular velocity of the forcing cycle. The creep function is given by (e.g., Efroimsky 2012a, 2012b)

Equation (B2)

where μ is the shear modulus, η is the viscosity, and 0 < α < 1 is the Andrade exponent. We assume α = 0.33 for ice (e.g., Rambaux et al. 2010). The relaxation time of the Andrade model is given by τA = η/μ.

We note that $\tilde{\mu }$ depends on a dimensionless parameter, τA ω, and can be divided into two regions. For the case of τA ω ≫ 1, the complex shear modulus is approximately given by

Equation (B3)

and $| \tilde{\mu }(\omega )| /\mu \simeq 1$, respectively. On the other hand, when τA ω ≪ 1, the complex shear modulus is approximately given by

Equation (B4)

and $| \tilde{\mu }(\omega )| /\mu \simeq {\tau }_{{\rm{A}}}\omega $, respectively.

B.2. Love Number

For a homogeneous spherical body, the tidal potential Love number, ${\tilde{k}}_{2}(\omega )$, is given by

Equation (B5)

where $g={ \mathcal G }M/{R}^{2}$ is the surface gravity, ${ \mathcal G }$ is the gravitational constant, and ρ and R are the density and radius of the body, respectively. We introduce the (dimensionless) effective rigidity, μeff, as follows (e.g., Goldreich & Sari 2009; Cheng et al. 2014):

Equation (B6)

and we found that μeff ∼ 102 for 1000 km sized TNOs. The imaginary part of the complex Love number is negative when ω is positive. When the angular velocity of the forcing cycle is negative, the complex Love number is given by

Equation (B7)

where $\overline{{\tilde{k}}_{2}(\omega )}$ is the complex conjugate of ${\tilde{k}}_{2}(\omega )$.

In the case of μeff ≫ 1, the tidal potential Love number is classified into three regions: (i) high-frequency region, (ii) intermediate region, and (iii) low-frequency regions (see also Efroimsky 2012b). We found that $\left|\mathrm{Im}\left[{\tilde{k}}_{2}(\omega )\right]\right|$ reaches a maximum at ${\tau }_{{\rm{A}}}\omega \simeq \pm {\mu }_{{\rm{eff}}}^{-1};$ the maximum value is $\left|\mathrm{Im}\left[{\tilde{k}}_{2}(\omega )\right]\right|\simeq 3/4$.

B.2.1. High-frequency Region

When τA ω ≫ 1, the absolute value and the imaginary part of the complex Love number, $\left|{\tilde{k}}_{2}(\omega )\right|$ and $\mathrm{Im}\left[{\tilde{k}}_{2}(\omega )\right]$, are approximately given by

Equation (B8)

Equation (B9)

The tidal quality factor, ${ \mathcal Q }(\omega )\equiv \left|\mathrm{Im}\left[{\tilde{k}}_{2}(\omega )\right]/{\tilde{k}}_{2}(\omega )\right|$, is given by

Equation (B10)

B.2.2. Intermediate Region

When ${{\mu }_{\mathrm{eff}}}^{-1}\ll {\tau }_{{\rm{A}}}\omega \ll 1$, $\left|{\tilde{k}}_{2}(\omega )\right|$, $\mathrm{Im}\left[{\tilde{k}}_{2}(\omega )\right]$, and ${ \mathcal Q }(\omega )$ are approximately given by

Equation (B11)

Equation (B12)

and

Equation (B13)

B.2.3. Low-frequency Region

When ${\tau }_{{\rm{A}}}\omega \ll {\mu }_{{\rm{eff}}}^{-1}$, $\left|{\tilde{k}}_{2}(\omega )\right|$, $\mathrm{Im}\left[{\tilde{k}}_{2}(\omega )\right]$, and ${ \mathcal Q }(\omega )$ are approximately given by

Equation (B14)

Equation (B15)

and

Equation (B16)

Appendix C: Cooling and Heating Rates of the Primary

In this study, we assumed that the interior of the primary is homogeneous for simplicity, then we calculated the time evolution of the internal temperature. We found that the temperature evolution is approximately given by the balance between conduction/convection cooling and decay heating. Here we show the cooling and heating rates of the primary.

The cooling rate of the primary due to conduction/convection is given by

Equation (C1)

Equation (C2)

Equation (C3)

where dTcond,G/dt and dTconv,G/dt are the cooling rates due to conduction and convection, respectively. Similarly, the heating rate of the primary due to decay the decay of long-lived radioactive elements is given by

Equation (C4)

As the specific heat and the thermal conductivity depend on the temperature, the cooling and heating rates also depend on temperature of the primary. We note that the decay heating rate is also a function of the time.

Figure C1 shows the cooling/heating rates of the primary as a function of the temperature, the reference viscosity, and the time. For the case of ηref = 1014 Pa s, the cooling rate is controlled by convection when TG > 206 K. We found that the cooling and heating terms balance at a certain temperature, and the equilibrium temperature is consistent with the temperature evolution of the primary shown in Appendix D. We also show that the equilibrium temperature depends strongly on ηref. For the case of ηref = 1010 Pa s, the equilibrium temperature is lower than 210 K even at t = 0. This low equilibrium temperature is also consistent with the temperature evolution shown in Figure D8(d).

Figure C1.

Figure C1. Cooling rate of the primary due to conduction/convection, dTcon,G/dt, and heating rate due to decay of long-lived radioactive elements, dTdec,G/dt.

Standard image High-resolution image

We also discuss the validity of the assumption of the fixed surface temperature. As shown in Appendix D, the internal temperature of the primary reaches the equilibrium temperature at t ≲ 1 Gyr. When we assume that the decay heating is balanced with the radiative cooling at the surface, the surface temperature of the primary, Tsurf, is given by the following equation:

Equation (C5)

where σSB is the Stefan–Boltzmann constant and TBG is the effective background temperature. Here we rewrite Tsurf as TBG + ΔT, then the difference between Tsurf and TBG is approximately given by

Equation (C6)

Equation (C7)

and ΔT is negligibly smaller than TBG. Thus we can assume Tsurf = TBG when the decay heating is balanced with the radiative cooling at the surface.

We acknowledge that our calculations of thermal evolution would not be accurate, especially for the early stage (i.e., t ≪ 1 Gyr). In the early stage, the temperature structure inside the primary may not reach the equilibrium, and the cooling rate should be modified when we calculate the internal temperature structure. As the thermal evolution in the early phase would be important for the eccentricity evolution of satellite systems (see Appendix D), we need to conduct the simulations of coupled thermal–orbital evolution considering the internal temperature structure in future studies.

Appendix D: Typical Results for Calculations of Tidal Evolution

In this appendix, we introduce typical tidal evolution pathways of satellite systems. We classify the outcome of tidal evolution into four types: 1:1 spin–orbit resonance (Type A; Figures D1 and D2), higher-order spin–orbit resonance (Type B; Figures D3 and D4), nonresonance (Type C; Figures D5 and D6), and collision with the primary (Type Z; Figure D7). We also show a typical evolution pathway with ηref = 1010 Pa s (Figure D8).

Figure D1.

Figure D1. Evolution of (a) semimajor axis, (b) eccentricity, (c) spin/orbital periods, and (d) temperatures of the primary and secondary for Model A. The initial condition of Model A is RX = 60 km, Tini = 200 K, and eini = 0.4.

Standard image High-resolution image
Figure D2.

Figure D2. Early phase of the evolution of (a) semimajor axis, (b) eccentricity, (c) spin/orbital periods, (d) temperatures of the primary and secondary, and (e) spin-to-orbit period ratio for Model A.

Standard image High-resolution image
Figure D3.

Figure D3. Evolution of (a) semimajor axis, (b) eccentricity, (c) spin/orbital periods, and (d) temperatures of the primary and secondary for Model B. The initial condition of Model B is RX = 60 km, Tini = 140 K, and eini = 0.4.

Standard image High-resolution image
Figure D4.

Figure D4. Early phase of the evolution of (a) semimajor axis, (b) eccentricity, (c) spin/orbital periods, (d) temperatures of the primary and secondary, and (e) spin-to-orbit period ratio for Model B.

Standard image High-resolution image
Figure D5.

Figure D5. Evolution of (a) semimajor axis, (b) eccentricity, (c) spin/orbital periods, and (d) temperatures of the primary and secondary for Model C. The initial condition of Model C is RX = 60 km, Tini = 120 K, and eini = 0.4.

Standard image High-resolution image
Figure D6.

Figure D6. Early phase of the evolution of (a) semimajor axis, (b) eccentricity, (c) spin/orbital periods, (d) temperatures of the primary and secondary, and (e) spin-to-orbit period ratio for Model C.

Standard image High-resolution image
Figure D7.

Figure D7. Evolution of (a) semimajor axis, (b) eccentricity, (c) spin/orbital periods, and (d) temperatures of the primary and secondary for Model Z. The initial condition of Model Z is RX = 60 km, Tini = 200 K, and eini = 0.1.

Standard image High-resolution image
Figure D8.

Figure D8. Evolution of (a) semimajor axis, (b) eccentricity, (c) spin/orbital periods, and (d) temperatures of the primary and secondary for a typical run with ηref = 1010 Pa s. The initial condition of this run is ηref = 1010 Pa s, RX = 60 km, Tini = 140 K, and eini = 0.4.

Standard image High-resolution image

D.1. Type A: 1:1 Spin–Orbit Resonance

Figure D1 shows a typical tidal evolution pathway of a satellite system resulting in Type A. The initial condition of this model is RX = 60 km, Tini = 200 K, and eini = 0.4 (see Figure 3(c)). The final semimajor axis and eccentricity are afin/RG = 26.0 and efin = 5.6 × 10−3, respectively; and the spin of the secondary is in 1:1 spin–orbit resonance at t = 4.5 Gyr.

The thermal evolution of both the primary and secondary hardly depends on their spin/orbital evolution. This is because their thermal evolution pathways are determined by the balance between decay heating and conduction/convection cooling.

Figure D2 shows the first 10 Myr of the tidal evolution of this case. Both the primary and secondary are initially captured into spin–orbit resonances. At t = 2.872 Myr, the eccentricity becomes e = 4.15 × 10−2, and the secondary is released from 5:2 spin–orbit resonance. Then, the primary is also released from 2:1 spin–orbit resonance at t = 2.875 Myr. The secondary is recaptured into 2:1 spin–orbit resonance at t = 2.897 Myr, and finally trapped in 1:1 resonance at t = 3.54 Myr.

D.2. Type B: Higher-order Spin–Orbit Resonance

Figure D3 shows a typical tidal evolution pathway of a satellite system resulting in Type B. The initial condition of this model is RX = 60 km, Tini = 140 K, and eini = 0.4 (see Figure 3(c)). The final semimajor axis and eccentricity are afin/RG = 29.7 and efin = 0.37, respectively, and the spin of the secondary is in 5:2 spin–orbit resonance at t = 4.5 Gyr.

The temperature of the 1000 km sized primary hardly depends on the initial temperature approximately 1 Gyr after the start of calculations (see Figures D1(d), D3(d), and D5(d)). This shows that the temperature of the primary reaches the equilibrium determined by the balance of decay heating and conduction/convection cooling at t ≲ 1 Gyr.

Figure D4 shows the first 300 Myr of the tidal evolution of this case. Both the primary and secondary are initially captured into spin–orbit resonances. The primary is released from 2:1 spin–orbit resonance at t = 113 Myr, at which time the eccentricity begins increasing. Conversely, 5:2 spin–orbit resonance of the secondary is stable in this case.

D.3. Type C: Nonresonance

Figure D5 shows a typical tidal evolution pathway of a satellite system resulting in Type C. The initial condition of this model is RX = 60 km, Tini = 120 K, and eini = 0.4 (see Figure 3(c)). The final semimajor axis and eccentricity are afin/RG = 28.0 and efin = 0.32, respectively, and the spin of the secondary is not in spin–orbit resonances at t = 4.5 Gyr.

Figure D6 shows the first 300 Myr of the tidal evolution of this case. The secondary is captured into 3:1 spin–orbit resonance at first, and then the primary is captured into 2:1 spin–orbit resonance. Both the primary and secondary are released from the spin–orbit resonance at t = 193.8 Myr, and the breaking of the spin–orbit resonance of the primary is prior to that of the secondary. The eccentricity starts to increase at this time, similar to the case shown in Figure D4. In Section 4.2, we show the conditions required to maintain spin–orbit resonances.

D.4. Type Z: Collision with the Primary

Figure D7 shows a typical tidal evolution pathway of a satellite system resulting in Type Z. The initial condition of this model is RX = 60 km, Tini = 200 K, and eini = 0.1 (see Figure 3(c)). The final semimajor axis and eccentricity are afin/RG = 1.1 and efin = 1.9 × 10−5, and the spin of the secondary is in 1:1 spin–orbit resonance at t = 56 kyr. In this case, the secondary collides with the primary at the end of the simulation. Or, in reality, the secondary is tidally disrupted and second-generation satellites/rings might be formed after the disruption event.

D.5. Undifferentiated Bodies Made of Soft Ice

Figure D8 shows a tidal evolution pathway of a satellite system for the case of ηref = 1010 Pa s. The initial condition of this model is RX = 60 km, Tini = 140 K, and eini = 0.4 (see Figure 8(d)). The final semilatus rectum is not very different from that of the standard case of ηref = 1014 Pa s (see Figures D1(d), D3(d), and D5(d)), because the temperature of Gonggong, TG, also depends on ηref. In the case of ηref = 1010 Pa s, TG is lower than that for the case of ηref = 1014 Pa s; and the viscosity of Gonggong, ηG = η(TG), is not very different between the two settings, ηref = 1010 Pa s and ηref = 1014 Pa s.

Footnotes

  • 4  

    The diameter of (90377) Sedna is 995 ± 80 km (Pál et al. 2012) and it has no known satellites. The others, namely (134340) Pluto, (136199) Eris, (136108) Haumea, (136472) Makemake, (225088) Gonggong, and (50000) Quaoar, have one or more satellites.

  • 5  

    Strictly speaking, the simplifying assumption of homogeneous bodies is a good approximation only for the convective case and for bodies without liquid layers (e.g., Bolmont et al. 2020). However, in this study, we apply Equation (25) for both convective and conductive cases. We should consider the internal temperature structure in future studies, although it will involve a high numerical cost.

Please wait… references are loading.
10.3847/1538-3881/ac1f91