The following article is Open access

A Past Episode of Rapid Tidal Evolution of Enceladus?

and

Published 2023 July 11 © 2023. The Author(s). Published by the American Astronomical Society.
, , Citation Matija Ćuk and Maryame El Moutamid 2023 Planet. Sci. J. 4 119 DOI 10.3847/PSJ/acde80

Download Article PDF
DownloadArticle ePub

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

2632-3338/4/7/119

Abstract

Saturn possesses a dynamically rich system containing numerous moons and impressive rings. Whether the rings of Saturn are much younger than the planet itself has been a long-standing open question; more recently a young age has been proposed for some moons. The recent detection of the fast orbital evolution of Rhea and Titan strongly suggests a highly frequency-dependent tidal response of Saturn, possibly through the excitation of inertial waves within the planet's convective envelope. Here we show that resonance locking to inertial waves cannot explain the dynamical structure of the Saturnian system or the current tidal heating of Enceladus. On the other hand, both the observations and our modeling results indicate that the system is not consistent with evolution under equilibrium tides. We propose that the system's architecture can best be explained by a relatively high "background" tidal response coupled with discrete resonant modes. In this view, only Titan may be in a true long-term resonance lock with a tidal mode of Saturn. Rhea is most likely currently experiencing a transient period of fast tidal evolution as it passes through a mode, rather than being locked into it. Assuming that Enceladus went through a temporary period of fast tidal evolution, we can reproduce its present resonance with Dione and satisfy other dynamical constraints. Additionally, we conclude that the long-term tidal response of Saturn to Tethys must be weaker than expected from frequency-independent tides, as already found by observations.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 4.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

Before the Cassini mission to Saturn (2004–2017) it was widely thought that the major moons of Saturn (Mimas and larger) are as old as the planet, and that the moons' orbital evolution is driven by equilibrium tides within Saturn. "Equilibrium tides" refers to classical tidal theory described by textbooks like Murray & Dermott (1999), in which there is little or no dependence of tidal dissipation on the orbital frequency of the perturbing moon. These two assumptions constrained the tidal quality factor Q which quantifies tidal dissipation within Saturn to Q > 18,000 (NB: lower Q means higher dissipation), or else Mimas would have been within the rings fewer than 4.5 Gyr ago (Murray & Dermott 1999).

The detection of thermal flux of 10–15 GW on Enceladus (Porco et al. 2006; Howett et al. 2011) challenged this estimate of the tidal evolution rate. Meyer & Wisdom (2007) showed that, assuming an equilibrium state, tidal heating of Enceladus through its orbital resonance with Dione produces (18,000/Q) × 1.1 GW. Their result implies that either the tidal Q of Saturn is an order of magnitude lower than 18,000, or the heating of Enceladus is not in equilibrium. The latter explanation of the observations was initially dominant, but the astrometric work of Lainey et al. (2012) suggested a migration rate of most major moons that is an order of magnitude faster than previously estimated. A notable implication of the results of Lainey et al. (2012) was that the tidal Q ≈ 1700, and the tidal heating of Enceladus could be in equilibrium.

Ćuk et al. (2016), assuming equilibrium-type tides and a constant tidal Q ≃ 1700 as found by Lainey et al. (2012), analyzed the orbital histories of the three largest moons interior to Titan: Tethys, Dione, and Rhea. Ćuk et al. (2016) found that their orbits are consistent with Dione and Rhea crossing their mutual 5:3 mean-motion resonance (MMR) in the past. Ćuk et al. (2016) also modeled past passage of Tethys and Dione through their 3:2 MMR and found that this event excites inclinations well above the observed values, implying that this resonance passage did not happen. Assuming Q = 1700, this relative dynamical age (i.e., Dione–Rhea 5:3 MMR was crossed in the past, but Tethys–Dione 3:2 MMR was not) translates to an absolute age of the system below about 100 Myr. While the age of Saturn's rings is hotly debated, this result is consistent with the estimate of the rings' age derived by Cuzzi & Estrada (1998). Ćuk et al. (2016) concluded that the rings and moons interior to Titan formed in a dynamical instability about 100 Myr ago, in which the previous generation of moons was disrupted in collisions, and then largely reaccreted into the observed satellites.

Some of the more recent findings are consistent with a young rings and satellite system. The relatively small mass (Iess et al. 2019) and rapid evolution (O'Donoghue et al. 2019) of the rings suggest a relatively young age, but other authors argue that ring pollutants are lost faster than ice, implying an older age (Crida et al. 2019). It appears that dominant past impactors in the Saturnian system are different from Kuiper Belt objects (Zahnle et al. 2003; Singer et al. 2019), and are possibly planetocentric (Ferguson et al. 2020, 2022a, 2022b). This would be consistent with, but would not require, a recent origin of the system. More recently, Wisdom et al. (2022) proposed an alternate proposal for a recent cataclysm that originates not in the inner system but in an instability between Titan and a past resonant moon (see Asphaug & Reufer 2013; Hamilton 2013); the full consequence of this scenario for the inner moons is still unclear.

Since the work done by Ćuk et al. (2016), analysis of the moons' motions using both Earth-based astrometry and Cassini data (Lainey et al. 2017, 2020; Jacobson 2022) have strongly suggested that the tidal evolution in the Saturnian system is not driven by equilibrium tides. Rhea has been found to migrate outward many times faster than predicted by equilibrium tides, with an orbital evolution timescale of $a/\dot{a}=6\,{Gyr}$. Additionally, Lainey et al. (2012) find that Titan is migrating with an 11 Gyr timescale, but that is disputed by Jacobson (2022), who find a >100 Gyr timescale for Titan's evolution. Both groups agree on the fast evolution of Rhea, ruling out equilibrium tides as the only form of tidal dissipation within Saturn.

The pattern of tidal evolution found by Lainey et al. (2020) matches the expectations of the resonant locking theory of Fuller et al. (2016). In this theory, Saturn's tidal response is exceptionally high at certain synodic frequencies, at which the tidal perturbations from the moons are resonant with internal oscillation modes of the planet. If the planet's structure were static, the moons would just quickly evolve through these frequencies without much consequence for their long-term evolution. However, due to changes in the resonant frequencies of the planet, the orbital locations at which a moon is resonant with the planet move outward, pushing the moons along faster than they would move through equilibrium tides (Fuller et al. 2016). If this is true, each moon is migrating due to a "resonance lock" with the planet's interior. Similar migration timescales for multiple moons imply that inertial waves within the planet are the type of oscillation driving the evolution. Lainey et al. (2020) suggest that the rate of tidal evolution for all moons at any time in their history may be given as $\dot{a}/a={(3{t}_{s})}^{-1}$, where ts is the age of the planet at any time.

Orbital evolution of all of Saturn's moons with a uniform relative migration rate rules out mutual MMR crossings; while approximately equal, but not identical, evolution rates would greatly extend the time between any MMR crossings. If this is the mechanism behind the evolution of the Saturnian system, the constraints on its age proposed by Ćuk et al. (2016) on the basis of MMRs do not apply. However, MMRs currently present in the system, as well as secular resonances, still need to be reconciled with this hypothesis of parallel orbital evolution.

Equilibrium tides that have previously been assumed to dominate the orbital evolution of Saturnian satellites have a strong dependence on orbital distance, $\dot{a}/a\propto {a}^{-6.5}$ (Murray & Dermott 1999). On the other hand, resonant lock produces orbital evolution that is either independent of orbital distance, $\dot{a}/a\ne f(a)$, in case of locking to inertial waves, or is faster for more distant satellites, $\dot{a}/a\propto {a}^{3/2}$, in case of locking to Saturn-frame modes (Fuller et al. 2016; Lainey et al. 2020). Therefore, it is likely that equilibrium tides would dominate the evolution in a zone closer to the planet, while resonance lock (if present) would dominate the evolution of more distant satellites. The exact boundary between these two regimes depends on the strength of the planet's tidal response and the rate at which the resonant peaks in the tidal response are shifting. Note that the equilibrium tidal evolution rate is directly proportional to the satellite's mass, while the resonance-lock evolution rate is independent of it, so the distance at which each of the mechanisms dominates will not be the same for moons of all sizes. Additionally, in the innermost zone close to Saturn's rings, ring-torque-driven evolution (Goldreich & Tremaine 1979) may dominate over all types of tidal evolution.

Here we will consider several specific dynamical features of the Saturnian system in the light of potential mechanisms of tidal evolution.

2. Constraints on Saturn's Tidal Response from Current Tidal Heating of Enceladus

Enceladus and Dione are currently locked in an MMR with argument 2λD λE ϖE , where λ and ϖ are the mean longitude and the longitude of pericenter, respectively (Murray & Dermott 1999), while subscripts D and E respectively refer to Dione and Enceladus. This resonance keeps the moons' orbital period ratio close to 1:2, and acting to increase the eccentricity of Enceladus over time. Satellite tides within Enceladus act in the opposite direction, damping Enceladus's eccentricity and releasing heat in the process. This tidal dissipation within Enceladus is thought to drive the observed geological activity on Enceladus (Porco et al. 2006).

The amount of heat observed on Enceladus (Howett et al. 2011), combined with a relatively low eccentricity (eE = 0.005), implies that Enceladus is very dissipative with tidal parameters given by Q/k2 ≈ 100 (or even smaller if there is more undetected heat), consistent with a fluid response to tides (Lainey et al. 2012). Given that this eccentricity would dampen in under 1 Myr, it is natural to assume that the eccentricity of Enceladus is continuously reexcited by the resonance with Dione, so that the eccentricity and the tidal heating are in a long-term equilibrium. In the context of equilibrium tides, this would require Q ≈ 1700 for Saturn (Meyer & Wisdom 2008; Lainey et al. 2012). The dynamics of the tidal heating of Enceladus in the resonance-lock scenario are yet to be fully modeled.

For the purposes of an MMR between two moons, there are several possible configurations of resonant modes. The simplest case is that only the inner moon is locked with a resonant mode, with the outer moon evolving only due to interaction with the inner. Alternatively, two moons could be locked to two converging modes. Convergent modes, however, are not compatible with a moon–moon resonance, as the inner moon would push the outer moon outward away from its mode, and we are back to the original case of only the inner moon evolving through resonant locking. A third possibility is simultaneous divergent resonant modes, but they would preclude capture into the moon–moon MMR, which requires convergent evolution (Murray & Dermott 1999). A fourth possibility is also conceivable: two modes evolving in parallel, with the inner moon experiencing stronger torque by being closer to the mode center, while the outer moon is subject to outward acceleration, both from the resonant mode and the perturbations of the inner moon. While this configuration is in principle consistent with an MMR, it is clear that it is sensitively dependent on the moons' locations relative to the resonant modes, and that the stability of this state would need to be investigated in greater detail. Therefore, here we will concentrate on the scenario in which only the inner moon is in resonance lock.

It can be shown that an MMR driven by resonance locking of the inner moon alone due to inertial modes is not consistent with the present heating rate of Enceladus (assuming equilibrium). Meyer & Wisdom (2007) show that the power of tidal heating of the inner moon of a resonant pair, assuming no tidal torque on the outer moon and all eccentricities being in equilibrium, is:

Equation (1)

where TE is the tidal torque on Enceladus, while L is the angular momentum, n is the mean motion, a is the semimajor axis, m is the mass, and G is the gravitational constant. Given that the moons are evolving in parallel due to MMR lock, the torque has a simple relation with the evolution rate (Meyer & Wisdom 2007):

Equation (2)

where ta is the evolution timescale. Substituting this back into Equation (1), and assuming that the angular momentum and energy of Dione are much bigger than those of Enceladus (accurate to about 10%), we get:

Equation (3)

Recognizing that ${{Gm}}_{S}/{a}_{D}^{3}={n}_{D}^{2}$, nE ≈ 2nD , and, for small eccentricities, ${L}_{D}={m}_{D}{a}_{D}^{2}{n}_{D}$, we get:

Equation (4)

If we assume ta = 9 Gyr, as proposed by Lainey et al. (2020) as being a typical value for orbital evolution due to resonance locking in the Saturnian system, H ≈ 180 GW, more than an order of magnitude in excess of the observed value (Howett et al. 2011). Therefore it appears that the observed resonance between Enceladus and Dione cannot be maintained by only Enceladus being locked to an inertial wave within Saturn.

On the other hand, if the resonance lock evolves uniformly in a reference frame rotating with Saturn, with the dominant semidiurnal tidal period changing by a constant absolute rate for all satellites (Fuller et al. 2016), the steady-state heating of Enceladus will be much lower. The Saturn-frame resonance lock would produce $a/\dot{a}\approx 200\,\mathrm{Gyr}$ for Enceladus, based on Titan's evolution timescale of 11 Gyr (Lainey et al. 2020) and $\dot{a}/a\propto {a}^{3/2}$ (Fuller et al. 2016). This rate of evolution gives us a steady-state heat flow of about 8 GW for Enceladus, close to but below the measured value (which is by itself a lower limit of the real flux). One caveat with this calculation is that we assumed no contribution from equilibrium tides acting on Enceladus and (more importantly) Dione, meaning that Saturn's "background" Q/k2 ≫ 30,000, or otherwise the tidal heating would be even lower.

If we follow recent results of Jacobson (2022) who find a possible resonant lock for Rhea but not Titan, the timescale for Enceladus's evolution through resonant locking (assuming divergent modes anchored at Rhea) could be in the 30–40 Gyr range, with the associated tidal heating in the 40–55 GW range. Models of Enceladus's ice shell suggest a heat loss rate in the range 25–40 GW (Hemingway & Mittal 2019), and the total tidal heating rate of Enceladus likely exceeds the measured value (possibly due to distributed tidal heating outside the South Polar Terrain), so there may not be a discrepancy between these predictions and the actual heat production rate.

If we assume equilibrium tides and a steady state, then a Q/k2 = 5000 can explain the observed heat flux of Enceladus, as suggested by Lainey et al. (2012). Note that a combination of inertial wave locking and equilibrium tides can also explain the observed flux, as long as the equilibrium tidal evolution rate of Dione is about 90% of the drift rate of the inertial mode to which Enceladus is locked. This complex (but possible) setup still gives us about the same "background" Q/k2 = 5000 as the assumption of pure equilibrium tides. The possibility of Enceladus and Dione being locked to two normal modes evolving almost in unison is also conceivable (and would result in lower equilibrium heat flows), but we assess it as less likely.

3. Constraints on Saturn's Tidal Response from Orbital Resonances

3.1. Past 2:1 MMR between the Horseshoe Moons and Enceladus?

Janus and Epimetheus are inner moons of Saturn that are currently in a horseshoe orbital configuration which undergoes a reversal every four years (Murray & Dermott 1999). Gravitational interactions with Saturn's rings make the width of the "horseshoe" decrease over time, and the timescale for this evolution is on the order of 107 yr (Lissauer 1985). For our purposes Janus and Epimetheus (and their presumed parent body) are interesting because they are expected to migrate very fast due to ring torques, potentially crossing resonances with other moons. Tajeddine et al. (2017) estimate that Janus migrates about 40 km Myr−1 due to ring torques, and the two moons' putative parent body (if they originated in a breakup) would have migrated even faster, at 50 km Myr−1. Janus and Epimetheus are about 1500 km exterior to the 2:1 MMR with Enceladus. Ignoring the tidal evolution of Enceladus for the moment, it appears that the precursor of Janus and Epimetheus should have crossed this resonance about 30 Myr ago, putting constraints on the age and evolution of the system.

We simulated this resonance crossing between coorbitals and Enceladus using the numerical integrator simpl, previously used by Ćuk et al. (2016). Enceladus was assumed to be in its present orbital resonance with Dione. At first we assumed that Janus and Epimetheus were in a horseshoe configuration before the resonance, and gave them their present eccentricities and inclinations. The left-hand panels in Figure 1 show a typical outcome of such simulations, in which the horseshoe configuration is disrupted and Janus and Epimetheus experience close encounters (our integrator does not test for collisions). This destabilization happens well before the core of the 2:1 resonance with Enceladus is reached, implying that the horseshoe is disrupted by near-resonant perturbations.

Figure 1.

Figure 1. Simulations of the 2:1 MMR crossing between the of coorbitals Janus and Epimetheus and Enceladus using simpl. Enceladus and Dione were assumed to be on their current orbits, while Janus and Epimetheus were shifted inward. In the left-hand side panels Janus and Epimetheus were in their current horseshoe configuration, while they were put in a tadpole (Trojan) configuration in the right-hand panels. The coorbitals' orbits are expanding due to an artificial acceleration meant to represent ring torques. We used k2/Q = 0.01 for Enceladus, and we ignored satellite tides for the coorbitals.

Standard image High-resolution image

Next we assumed that the two moons were in corbitals before the resonance, but in a Trojan or "tadpole" configuration (Murray & Dermott 1999); typical results are shown in the right-hand side panels of Figure 1. Interestingly, near-resonant perturbations always convert the tadpole configuration into a horseshoe, which is stable from there on. However, the moons are then captured into a 2:1 resonance with Enceladus, in which their eccentricities grew over time while the semimajor axis is locked to that of Enceladus. This eccentricity growth is largely unaffected by any tidal damping, as the small sizes and demonstrably rigid natures of the irregularly shaped Janus and Epimetheus do not allow for significant tidal deformation or dissipation over the 107 yr timescales in which we are interested. We expect the resonance eventually to break and Janus and Epimetheus to collide. As we expect the moons to have large eccentricities at the time of the collision while their semimajor axes are initially within resonance with Enceladus, the resulting debris must reaccrete significantly interior to the resonance (due to angular momentum conservation).

We also ran a simulation assuming that Janus and Epimetheus were contained within one body during the resonance crossing with Enceladus (which may have subsequently broken up). Our simulations of the 2:1 resonance crossing with Enceladus show that the precursor is always captured in the resonance (Figure 2). Note that we assumed the current orbital eccentricity of Janus for the progenitor, and the present-day configuration of the Enceladus–Dione resonance. In most simulations, proto-Janus is captured into a stable corotation resonance in which the eccentricity of Enceladus is slightly higher than equilibrium eccentricity in the resonance with Dione (left-hand panels in Figure 2). In some cases this resonance is broken and proto-Janus is captured into the eE (i.e., Lindblad) subresonance in which the eccentricity of proto-Janus grows over time, like we saw for the initial tadpole configuration. Given the large orbital precession rates of the innermost moons, all resonances are well separated and there is no obvious way of breaking this lock purely through dynamics. However at eccentricities in the 0.05–0.1 range proto-Janus would cross the orbits of Prometheus and Pandora (if they were present at that epoch) or the outer edge of the rings. Resulting collisions would presumably break the resonant lock between proto-Janus and Enceladus. Once again, angular momentum conservation dictates that any reaccretion of proto-Janus's debris happens interior to the resonance with Enceladus.

Figure 2.

Figure 2. Simulations of the 2:1 MMR crossing between the parent body of the coorbitals of Janus and Epimetheus and Enceladus using simpl. The coorbitals' parent body was given their combined mass and was put on an orbit interior to Enceladus resonance with the current e and i of Janus. The progenitor's orbit is expanding due to an artificial acceleration approximating ring torques. The two simulations differ slightly in initial conditions. We used k2/Q = 0.01 for Enceladus, and we ignored satellite tides for the coorbitals' precursor.

Standard image High-resolution image

If we assume that the coorbitals were produced in the breakup of a single progenitor, we can use their orbits to constrain the orbit of this progenitor. Using the current orbits of the coorbitals, and under the assumption of conservation of momentum at separation, we find that the prebreakup parent body of Janus and Enceladus had an eccentricity e ≤ 0.006, much less than expected from the past capture into the 2:1 resonance with Enceladus. Eccentricity can be erased through reaccretion, but as we discuss above, disruption and reaccretion move the resulting new moons interior to the resonance with Enceladus. Therefore, either the tidal evolution of Enceladus has to be comparable with the orbital evolution of the horseshoe pair, or the inner Saturnian system has to be rather young (a combination of these two factors is also possible).

If the resonant pair Enceladus and Dione are evolving through equilibrium tides with Q/k2 = 5000 for Saturn (Lainey et al. 2012), their joint orbital evolution timescale would be $a/\dot{a}\approx 10\,\mathrm{Gyr}$. Enceladus being locked to an inertial mode would produce the same timescale (even if the amount of tidal heating would be very different, as all of Dione's orbital evolution would now be due to Enceladus). This rate of evolution would only push the resonance with the Janus–Epimetheus precursor back to 40 Myr. Locking of Enceladus to a Saturn-frame mode would lead to $a/\dot{a}\simeq 130\,\mathrm{Gyr}$, which would produce negligible migration compared to the ring-torque-driven evolution of the coorbitals' precursor, keeping the apparent age of the resonant encounter at 30 Myr ago. Enceladus's rate of evolution through equilibrium tides (but not resonance locking) depends on its MMR with Dione. Before the resonance Enceladus would have had an equilibrium tidal evolution timescale $a/\dot{a}=5\,\mathrm{Gyr}$ (assuming Q/k2 = 5000 for Saturn), so depending on the age of the Enceladus–Dione resonance, the progenitor–Enceladus resonance would have happened between 40 and 75 Myr ago.

Lainey et al. (2020) propose that a model in which equilibrium tides are weak and the moons' orbits evolve solely through resonance locking may result in a long-term stable system. In this model, the moons' orbits are not expected to converge and enter MMRs with each other. In this way dynamical excitation and possible destabilization through MMRs are avoided. The simulations discussed here show that the inclusion of ring torques introduces relatively rapid convergence of the orbits of some of the inner moons. A young age (<50 Myr), at least for the Janus–Epimetheus parent body (JEPB) if not for other moons, appears inevitable. If we assume the presence of both equilibrium tides and resonant tidal response (Section 5.3) the actual maximum age for the JEPB may be somewhat older than 75 Myr but difficult to calculate precisely. Also, it appears that the JEPB formed close to its present location, and that the idea that Janus and Epimetheus evolved from the rings smoothly to their present positions (Crida & Charnoz 2012) may not be tenable given the apparent impossibility of the pair (or their parent body) crossing the 2:1 resonance with Enceladus.

3.2. Establishment of the Current Mimas–Tethys 4:2 Resonance

Mimas and Tethys are currently in a 4:2 inclination-type resonance that involves the inclinations of both moons (the resonant argument is 4λΘ − 2λM − ΩM − ΩΘ), where the subscripts M and Θ refer to Mimas and Tethys, respectively. A large libration amplitude of the resonance has in the past been used to suggest that the resonance capture was a low-probability event (Allan 1969; Sinclair 1972). Our work so far (under the assumption of equilibrium tides) suggests that there are at least two distinct dynamical pathways to resonance capture, both of which require a later increase of the resonance argument libration width due to a third body.

If the initial inclination of Mimas was very low (iM ≃ 0fdg001), the lack of capture into the preceding iM 2 subresonance and capture into iM iΘ subresonance are highly likely events (assuming iΘ ≈ 1° before the resonance). The right-hand panels in Figure 3 shows a typical capture into the Mimas–Tethys resonance assuming equilibrium tides with Q/k2 = 4000 for Saturn. This rate of tidal evolution implies that the current Mimas–Tethys resonance is only 20–25 Myr old, consistent with the relatively young proposed age of the system (Ćuk et al. 2016). The physical reason behind the lack of capture into the iM 2 subresonance appears to be the nonadiabatic nature of the resonance crossing, as proposed by Sinclair (1972). While the tidal evolution is smooth and slow by most criteria, the extremely narrow width of the iM 2 subresonance at very low inclinations seems to allow for nonresonant crossing.

Figure 3.

Figure 3. Simulations of the 4:2 MMR crossing and capture between Mimas and Tethys. The resonant argument plotted in bottom panels is 4λΘ − 2λM − ΩM − ΩΘ. In the left-hand simulation Mimas was initially assumed to have a very low inclination, while in the right-hand simulation we assumed an initial iM = 0fdg1 for Mimas. We assumed equilibrium tidal evolution with a frequency-independent Q/k2 = 4000 for Saturn, Q/k2 = 105 for Mimas, and Q/k2 = 104 for Tethys. The first jump in inclination is due to the i2 M resonance.

Standard image High-resolution image

Alternatively, if we assume a preresonant inclination of iM = 0fdg1 for Mimas, there is a high probability (about 70%–80%) of missing capture into the iM 2 subresonance, and a similarly high probability of capture into the iM iΘ subresonance. In this case the resulting libration amplitude of the resonant argument is about 60°, still short of the observed one but much larger than in the low-inclination limit. In this regime the capture into iM 2 is unlikely because of the large initial inclination (as is common for resonant encounters; Murray & Dermott 1999), while the capture into iM iΘ subresonance is somewhat more likely due to large iΘ (also, a negative kick to the iM when crossing iM 2 also makes the next capture more likely).

Figure 3 shows two of the many simulations we did to study capture into the 4:2 Mimas–Tethys resonance in isolation, meaning that we did not simulate the system's evolution before and after this encounter. Initially we expected the two subresonances of the 4:2 MMR (i2 M and iM iΘ) to be the only relevant inclination-type terms of the 2:1 Mimas–Tethys commensurability. Later simulations of the comprehensive recent dynamical history of the system have revealed additional relevant inclination-type resonances. In particular, a third-order harmonic of the 2:1 Mimas–Tethys resonance with the argument 2λΘλM + ϖM − ΩM − ΩΘ is surprisingly strong, and can lead to capture for low initial inclinations of Mimas (see Figure 4). This subresonance (which we designate eM iM iΘ) is encountered well before the iM 2 and iM iΘ subresonances shown in Figure 3. The reason for the unexpected strength of the eM iM iΘ resonant term is the high inclination of Tethys and high eccentricity of Mimas, which we assumed both predate these moons' 4:2 MMR. These assumptions were based on a lack of available mechanisms of exciting eM after the capture into the current 4:2 Mimas–Tethys resonance, and the necessity of large iΘ for the capture to occur. We find that an initial iM ≈ 0fdg1 is necessary in order to avoid capture into the eM iM iΘ subresonance. Therefore, while both low- and high-inclination routes to capture into the observed iM iΘ resonance are possible in isolation, additional constraints from the eM iM iΘ crossing make the high-inclination path the only likely capture mechanism.

Figure 4.

Figure 4. Simulations of the crossing and capture into a third-order subresonance of the Mimas–Tethys 2:1 MMR. The inclination of Mimas is plotted in four different simulations which started with eΘ = 0.002 and iΘ = 1° for Tethys, and eM = 0.02 and a range of initial inclinations for Mimas. The argument of this resonance is 2λΘλM + ϖM − ΩM − ΩΘ. The satellites' k2 and Q values are the same as in Figure 3.

Standard image High-resolution image

The captures into resonance shown in Figure 3 result in a libration amplitude of the resonant argument that is smaller than observed. In general, we find that that the libration amplitude of the Mimas–Tethys resonance is highly vulnerable to perturbations from other moons (including Enceladus, Section 4.2), so this resonance's libration amplitude is less indicative of the capture mechanism than previously thought. We note that the difference between the postcapture libration amplitude and the observed value is less pronounced in our preferred high-inclination route to Mimas–Tethys 4:2 MMR compared to the low-inclination route. Clearly more detailed study of the system's very recent evolution, accounting for the influence of all the moons (including Tethys's small Trojan companions), is necessary to understand better the evolution of the libration amplitude of the Mimas–Tethys resonance.

All hypotheses of the origin of the Mimas–Tethys resonance must include a convergent evolution of these two moons. This convergent evolution is expected in the case of equilibrium tides, but interestingly only in the "constant Q" rather than the "constant time-lag" model (as the synodic period of Mimas is relatively long). In the case of evolution by both moons being locked to inertial waves there should in principle be no relative orbital evolution, either convergent or divergent, while resonance locking to Saturn-frame modes would result in divergent evolution. If only Mimas is in resonant lock, then a resonance would form and should be about 20 Myr old (in the case of inertial waves) or about 300 Myr old (in case of locking to Saturn-frame modes). Locking of Mimas to a Saturn-frame mode would require a "background" Q ≥ 100,000, as faster equilibrium tidal migration would allow Mimas to "outrun" the mode.

Another common assumption in all scenarios on the origin of the Mimas–Tethys resonance requires Tethys to have a prior inclination of about iΘ ≈ 1°. This relatively large inclination requires some kind of past dynamical interaction with other moons, most likely a resonance. Ćuk et al. (2016) have found this inclination to be a plausible consequence of Dione–Rhea 5:3 resonance crossing, followed by the (associated) Tethys–Dione secular resonance. We note that this scenario requires convergent orbital evolution of Dione and Rhea, at least in the past (see Section 3.3).

3.3. Observed Acceleration of Rhea and the Past 5:3 Dione–Rhea Resonance

The orbit of Rhea is directly observed to evolve on a 6 Gyr timescale (Lainey et al. 2017, 2020; Jacobson 2022). The rate of orbital evolution of Titan may be similarly fast (Lainey et al. 2020), but this is still disputed (Jacobson 2022). Observational results for Rhea are clearly not consistent with the expectations based on equilibrium tides. Locking to resonant modes within Saturn has been proposed as an explanation for these observations (Fuller et al. 2016), with the apparent rate of the moon's orbital evolution being determined by the drift in frequency of the resonant peaks in the tidal response of the planet. After Rhea's fast evolution was first observed (Lainey et al. 2017), the assumption that the peaks in the tidal response have a constant absolute frequency drift in Saturn's rotating frame (Fuller et al. 2016) implied that Titan's own drift timescale should be $a/\dot{a}\lt 2\,\mathrm{Gyr}$. The observed tidal evolution rate of Titan of 11 Gyr (Lainey et al. 2020) would imply that the moons could not both be in resonant lock, or that the resonant mode drift is not constant in the rotating frame. Locking to inertial waves which drift together in an inertial reference frame was proposed instead (Lainey et al. 2020), which should result in the same tidal evolution timescale for all moons in resonant lock. The difference between the apparent evolution timescales for Rhea and Titan reported by Lainey et al. (2020) seems to suggest that moons under resonant lock can still have converging orbits, which would challenge the big picture of Lainey et al. (2020), in which moons never cross mutual resonances. On the other hand, if Titan has a much slower orbital evolution consistent with equilibrium tides (as found by Jacobson 2022), then we can be sure that not all moons are locked to resonant modes. Clearly the dynamics of the system is more complex than predicted in any one simple model.

The observed fast evolution of Rhea is at odds with the hypothesis that Dione and Rhea crossed their mutual 5:3 MMR. The Dione–Rhea 5:3 MMR crossing was proposed by Ćuk et al. (2016) as a way of producing the current inclinations of Tethys and Rhea. The inclination of Tethys is particularly significant (iΘ ≈ 1°) and must have predated Tethys's current resonance with Mimas. As Ćuk et al. (2016) show, the Dione–Rhea 5:3 resonance is usually followed by a secular resonance between Tethys and Dione during which Dione "passes" its eccentricity and inclination to Tethys. Therefore there is a compelling reason to think that a past Dione–Rhea 5:3 MMR crossing did take place.

The orbital evolution of Rhea therefore presents a conundrum. Its rate is much too fast for equilibrium tides (assuming Q/k2 = 5000), but it is also too fast for resonance locking (assuming that Titan is currently in resonance lock; Lainey et al. 2020). A Saturn-frame mode resonant lock would have $a/\dot{a}\approx 40\,\mathrm{Gyr}$, while a resonance lock with inertial modes would predict $a/\dot{a}=11\,\mathrm{Gyr}$, same as Titan. Furthermore, the apparent past crossing of the Dione–Rhea resonance implies that this fast evolution of Rhea is a very recent phenomenon (i.e., over the past several million years or so). Assuming all of these constraints are correct, the simple solution is that Rhea is not locked to a mode, but is currently crossing a resonant mode. This is a possible situation if a moon (through equilibrium tides) evolves faster than a mode rather than the other way around, making a resonant lock impossible but producing occasional bursts of fast orbital evolution. If we assume that Titan is currently evolving with a timescale of 100 Gyr (consistent with equilibrium tides; Jacobson 2022), then it is possible that resonance locking may not be feasible at all. The argument here is that Titan would have certainly been captured in a resonant lock if resonant modes moving faster than Titan's orbit evolve due to equilibrium tides. In principle, we can also envision resonant modes moving to higher frequencies over time, opposite the direction of tidal evolution. In this case, all accelerated tidal evolution would always be episodic, being driven by the moons overtaking resonant modes.

In the rest of the paper, we will examine how a combination of equilibrium tides and passage through resonant modes can explain other features of the Saturnian system, namely the establishment of the Enceladus–Dione MMR.

4. Enceladus–Dione 2:1 Resonance: Equilibrium Tides

In the previous sections we concluded that orbital evolution through equilibrium tides (with Q/k2 = 5000) is able to explain the present heating of Enceladus (Meyer & Wisdom 2007) and the capture of Mimas and Tethys into their current resonance with a high probability, given appropriate initial conditions. We also find that orbital evolution of Enceladus through equilibrium tides allows for an age of the Janus–Epimetheus pair (or their parent body) older than 40 Myr, unlike Enceladus's evolution through a resonant lock. However, it has not been explored in the prior literature whether equilibrium tides can establish the current Enceladus–Dione 2:1 MMR, as opposed to maintaining it. Study of this resonant capture must be done through numerical integrations, as Ćuk & El Moutamid (2022) have shown that practically all two-body resonances in the Saturnian system harbor three-body resonances that are very hard to model analytically, in addition to many two-body subresonances.

4.1. Enceladus–Dione 4:2 Inclination Resonances

The Enceladus–Dione 2:1 MMR contains two first-order subresonances, four more second-order ones, several important three-body resonances, as well as numerous two-body third-order subresonances. The currently occupied eE subresonance is one of the last to be encountered as the orbits of Enceladus and Dione converge. We ran a number of simulations of the Enceladus–Dione 2:1 MMR encounter assuming equilibrium tides and Q/k2 = 4000 for Saturn. Typically, Enceladus was captured into the 4:2 iE 2 resonances in a large majority of cases where Enceladus and Dione encounter this resonance under equilibrium tides. The left-hand panels in Figure 5 plot the result of a typical simulation that features capture into the Enceladus–Dione 4:2 iE 2 resonance. Enceladus's inclination grows to about 1°, at which point secondary resonances break the inclination resonance. In this particular simulation all inner moons (from Mimas to Rhea) were assumed to have initially equatorial orbits, but the result is the same when Tethys was assumed to have been already inclined. In all remaining simulations of equilibrium tide Enceladus–Dione 4:2 MMR crossing Enceladus is caught into a more complex inclination resonance, also affecting the inclination of Dione or Tethys. In all cases Enceladus acquires an inclination on the order of a degree, two orders of magnitude above the observed one.

Figure 5.

Figure 5. Simulations of the 4:2 MMR crossing and capture between Enceladus and Dione. The left-hand panels show capture into the pure Enceladus inclination resonance iE 2, the middle panels in Figure 5 show capture into the semi-secular three-body resonance involving Tethys with argument 4λD − 2λE − ΩE − ΩΘ, and the right-hand panels show capture into the mixed two-body 4:2 iE iD subresonance. We used Q/k2 = 104 for all three moons.

Standard image High-resolution image

The near certainty of capture into one of the inclination subresonances found in our simulations strongly suggests that the Enceladus–Dione 2:1 MMR was not assembled through equilibrium tides. This is in contrast to what we found for the Mimas–Tethys 4:2 resonance (Section 3), which can be accounted for using equilibrium tides (except for the present libration amplitude, which is easily excited by a third body). The high probability of capture into the Enceladus–Dione 4:2 iE 2 resonance, but not Mimas–Tethys 4:2 iM 2 resonance is rather surprising but holds in large numbers of numerical simulations. We suspect that the lack of capture into the Mimas–Tethys 4:2 iM 2 resonance with a low initial iM is due to the mechanism originally proposed by Sinclair (1972), which states that in cases of small free inclinations and fast orbital precession even a relatively slow resonant encounter may not be adiabatic, i.e., a second-order resonance may not be able to dominate the motion of the ascending node against oblateness-driven precession. Note that this explanation did not hold in numerical simulations using Q ≈ 18,000 for Saturn (Sinclair 1974), but appears to explain the behavior at the ten times faster tidal evolution used in this study.

4.2. Enceladus–Tethys 11:8 Resonance

In the course of our work on the past dynamics of the Saturnian system, we discovered that Enceladus–Tethys 11:8 resonance offers unexpected constraints on the past orbital evolution of those two moons. In the equilibrium tide paradigm, this resonance would have been crossed 10–15 Myr ago. Since Tethys's predicted equilibrium tidal evolution is faster than that of Enceladus (they are the only neighboring pair of major Saturnian moons on diverging orbits, as Tethys's six times larger mass boosts its tidal evolution), this resonance should have been crossed divergently without a chance of capture. Despite the third order of this subresonance, the large eccentricity of Enceladus and inclination of Tethys make this resonant term surprisingly relevant. In particular, we find that the Enceladus–Tethys 11:8 MMR crossing usually both excited the inclination of Enceladus beyond the observed value, and disrupted the Mimas–Tethys 4:2 MMR (Figure 6).

Figure 6.

Figure 6. Left-hand side: a simulation of recent evolution of the inner moons assuming frequency-independent equilibrium tides with Q/k2 = 4000 for Saturn. Initial conditions are set at about 30 Myr ago, and we assume that the Enceladus–Dione 2:1 MMR (2λD λE ϖE ) was already established and in equilibrium. Mimas and Tethys are captured into their 4:2 MMR (4λΘ − 2λM − ΩM − ΩΘ) at 15 Myr, but this resonance is prematurely broken by the interaction through the 11:8 Enceladus–Tethys MMR, while Enceladus acquires a significant inclination. Right-hand side: a very similar simulation assuming that tidal dissipation at Tethys's frequency is only 70% of that for the other moons. In this simulation Enceladus suffers inclination excitation before the beginning of the Mimas–Tethys 4:2 MMR, confirming that the (already inclined) Tethys is the relevant perturber, rather than (still noninclined) Mimas. We used Q/k2 = 105 for Mimas, Q/k2 = 102 for Enceladus, and Q/k2 = 104 for Tethys and Dione.

Standard image High-resolution image

These results decisively argue that the current system could not have resulted by evolution through equilibrium tides. Of course, observations of the tidal acceleration on Rhea (Lainey et al. 2017, 2020; Jacobson 2022) already established that Saturn's tidal response is dynamic, and that Rhea and/or Titan may be in a resonance lock. However, as we demonstrated in previous sections, the observed resonances between the inner moons appear to suggest convergent tidal evolution consistent with equilibrium tides. The apparent lack of past crossing of the Enceladus–Tethys 11:8 MMR is the first evidence from mutual resonances that requires frequency-dependent tides to have acted in the past.

In general, the Enceladus–Tethys 11:8 MMR could have been avoided through slower orbital evolution of Tethys or faster migration of Enceladus. We explored the former possibility assuming that the tidal response of Saturn is frequency dependent but does not contain resonant modes. This would be consistent with the findings of Lainey et al. (2020) and Jacobson (2022), who do find a slower orbital evolution of Tethys relative to the other inner moons than expected from equilibrium tides (however, the uncertainties are still large). The current tidal evolution rate of Enceladus is constrained by its tidal heating (Section 2), so a slower evolution of Tethys is necessary to avoid a recent passage through the Enceladus–Tethys 11:8 MMR. However, we have just seen that the equilibrium tides acting on Enceladus are still unable to reproduce the Enceladus–Dione 2:1 MMR (Section 4.1). Therefore it appears that the most likely solutions to the problem posed by the Enceladus–Tethys 11:8 MMR require both that (1) the tidal response at Tethys's frequency is currently weaker than that of Enceladus, and (2) that the migration of Enceladus was much faster in the past. We explore this possibility in the next section.

5. Enceladus–Dione 2:1 Resonance: Short-Lived Enhanced Tidal Evolution

In this section we present simulations of Enceladus and Dione encountering their 2:1 MMR while Enceladus is migrating much faster than expected by equilibrium tides. This is enabled in the numerical integrator simpl by adding a simple factor multiplying the tidal acceleration of each moon. For moons evolving under equilibrium tides this factor is simply 1, and it is larger for accelerated evolution. We can use this modification of equilibrium tides as our hypothesis is that Enceladus (and later Rhea) are simply passing through frequency bands that have a very high tidal response, rather than being locked to these resonant modes. Resonant locking cannot be modeled this way, as moons locked to resonant modes tend to have a constant migration rate but do not experience the constant tidal Q of Saturn (and the apparent tidal Q will change if the moon in question enters a resonance with an exterior satellite).

5.1. Inclination-type Subresonances of the Enceladus–Dione 4:2 MMR

It is a well-known phenomenon that the probability of MMR capture declines as the resonance is encountered more rapidly (Murray & Dermott 1999). In this subsection we explore the hypothesis that an episode of fast tidal evolution may have enabled Enceladus to cross the inclination resonances with Dione without capture. The results from a number of simulations are plotted in Figure 7. We ran 20 different simulations that had Saturn's Q/k2 = 4000 for all moons except Enceladus, for which the response was enhanced five times; six resulted in capture into an inclination-type resonance and 14 experienced a kick in inclination. In 20 simulations where Enceladus evolved 10 times faster than expected, none experienced capture. The postpassage inclination of Enceladus was on average lower in the 10× accelerated simulations (typically iE = 0fdg01−0fdg02) than in the 5× accelerated ones (typically iE = 0fdg02−0fdg04).

Figure 7.

Figure 7. Left-hand side: the results of 20 simulations (plotted together) of the crossing of the inclination subresonances of the Enceladus–Dione 4:2 MMR with Enceladus's orbit evolving at a 5× tidal rate. Right-hand side: 20 simulations of the same MMR crossing, now assuming Enceladus migrating at 10× the tidal rate. All the moons had their current inclinations and eccentricities at the start of the simulation, and each simulation had a different argument of pericenter of Enceladus, which led to a spread in the initial resonant arguments. The satellites' tidal parameters were the same as in Figure 5.

Standard image High-resolution image

Our choice of current inclinations and eccentricities for the initial conditions is somewhat arbitrary. While its present inclination is small, Enceladus presumably formed with close to zero inclination and this initial value may affect the postresonant distribution of inclinations. Later simulations (shown in Section 5.3) were started with a lower inclination of Enceladus but the outcomes were very similar. As seen in Figure 7, the three-body resonance with Tethys (middle of the three jumps) is a minor contributor to Enceladus's final inclination, so our assumption that Tethys was already inclined at this epoch is not crucial. The exact inclination of Dione may effect the results somewhat (as the third jump is the iE iD subresonance), but we see no strong reason to expect a different inclination for Dione at the time when 2:1 resonance with Enceladus was established.

5.2. Eccentricity-type Subresonances of the Enceladus–Dione 4:2 MMR

After the inclination-type subresonances have been crossed, Enceladus encounters a number of eccentricity-type subresonances of the Enceladus–Dione 2:1 MMR. Based on the results shown in Figure 7, we enhanced Enceladus's tidal evolution by a factor of 10, while the tidal evolutions of the other moons correspond to the frequency-independent Q/k2 = 4000 for Saturn. The choice of initial eccentricities requires making assumptions not only about past resonances but also about the rate of eccentricity damping by the moons.

In our first batch of simulations we used the current eccentricities for most moons, except that Enceladus was given a low initial eE < 0.001 while Dione was given eD = 0.004, about twice the current value (to account for the expected resonance kick and subsequent damping). Out of the first group of simulations using these initial conditions, the majority of runs resulted in Enceladus being captured into three-body resonances with Rhea or Tethys (Figure 8, first and third rows), and the rest resulted in the eD subresonance, with no simulations reaching the current eE subresonance (which is last in order of encounters).

Figure 8.

Figure 8. Example simulations of captures into various eccentricity subresonances of the Enceladus–Dione 2:1 MMR. The left-hand panels show the eccentricities of the relevant moons: Enceladus (dark green), Tethys (red), Dione (magenta), and Rhea (cyan), while the right-hand panels plot resonant arguments which are labeled on the corresponding left-hand plots. In the top three rows we assumed that Tethys and Rhea had their current low eccentricities, and in the bottom two rows we assumed their eccentricities were higher. Enceladus's migration rate was 10× tidal, and its initial eccentricity was assumed to be very low. We conclude that the low eccentricities of Tethys and Rhea lead to capture into first-order three-body resonances, while the low e of Enceladus leads to capture into second-order resonances (two- and three- body). These simulations are a subset of those shown on the right-hand side of Figure 7.

Standard image High-resolution image

Captures into three-body resonances, which were understandably missed in past semianalytical models (Meyer & Wisdom 2008) demonstrate once again the importance of such resonances in the Saturnian system (Ćuk & El Moutamid 2022). These captures may come as a surprise given the high rate of orbital migration of Enceladus, but we should recognize that these are first-order resonances, unlike the inclination-type resonances shown in Figure 7, which are necessarily of the second order. This preponderance of three-body resonance captures is in part a consequence of us starting the simulations with Tethys and Rhea on very-low-eccentricity orbits. However, these were not arbitrary choices, as these two moons currently do have almost circular orbits. The current low eccentricities of both Tethys and Rhea argue against these three-body resonance captures happening in the recent past, as damping the eccentricities acquired in Figure 8 over realistic timescales may be difficult, especially for Rhea. We therefore prefer these moons having relatively small eccentricities at this epoch and avoiding capture into the three-body resonances, rather than being captured and acquiring large eccentricities (on the order of 0.01 or more).

In the second batch of simulations, we gave both Tethys and Rhea eccentricities of 0.002, which we found (through trial and error) to be largely sufficient to avoid capture into three-body resonances. In the second group of simulations, Enceladus still never reaches the current eE resonance, but in the majority of cases becomes captured in a second-order resonance which also includes the eccentricity of either Titan or Dione (Figure 8, rows four and five, respectively). While the mixed resonance eE eD has been studied before (Meyer & Wisdom 2008), the three-body resonance with Titan is new. In both of these cases we cannot avoid capture into these subresonances by changing the eccentricity of Dione or Rhea. A less-eccentric Dione would make capture into eE eD resonance less likely, but would increase the fraction of captures into the first-order eD subresonance (this fraction is already 20%–30% for initial eD = 0.004) and may conflict with Dione's current eccentricity. In the case of Titan, its eccentricity does not get affected by the resonance and we must use values very close to the current one.

The simplest way to avoid capture into the abovementioned second-order eccentricity subresonances is to give Enceladus a sizable initial eccentricity. The implication is that at the time of capture into the Enceladus–Dione 2:1 MMR, not only was an accelerated orbital evolution of Enceladus necessary (to avoid capture into the inclination-type resonances), but all inner moons had to have somewhat eccentric orbits (to avoid capture into the eccentricity-type resonances). In the next section we will put together a possible evolution path that satisfies the constraints presented in this and previous sections, and the implications of the moons' previously excited eccentricities will be discussed in Section 6.

5.3. Comprehensive Model of the System's Recent Evolution

In this section we attempt to construct a numerical model of the recent (last 20 Myr) evolution of the Saturnian satellite system. We are using all of the constraints that were established in previous sections. Enceladus and Dione are assumed to experience tidal acceleration with Q/k2 = 4000 at the present epoch, 3 in agreement with the hypothesis of equilibrium tidal heating of Enceladus (Section 2; Meyer & Wisdom 2007; Howett et al. 2011; Lainey et al. 2012). Mimas is assumed to have the same tidal evolution rate, both because of a lack of other constraints, and as it is consistent with Mimas encountering and being captured into 4:2 resonance with Tethys (Section 3.2). Need to avoid the 11:8 Enceladus–Tethys resonance, as well as the (admittedly uncertain) results of Lainey et al. (2020) and Jacobson (2022), suggested that the tidal acceleration of Tethys is much weaker than that expected from equilibrium tides. Finally, and most radically, we assumed that Enceladus went through a stage of very rapid tidal evolution due to crossing of (but not locking to) a resonant mode. This was done to avoid capture into Enceladus–Dione 4:2 inclination resonances (Section 4.1), as well as to avoid a recent crossing of Enceladus's 1:2 resonance with the horseshoe moons (or their progenitor; Section 3.1). Additional assumptions include low initial tidal dissipation within Enceladus until it experiences strong tidal heating, at which point Enceladus switches to a strong tidal response (this is done to approximate melting of the interior). Due to the nature of the integrator where the tidal response is treated as a constant parameter, we modeled both the transitions in the tidal acceleration of Enceladus and the changes in its own tidal response as abrupt events, implemented manually.

Table 1 shows the initial conditions for two sets of simulations of the integrated evolution of the inner moons (sets differ only by initial iΘ). In both cases, Enceladus was assumed to be experiencing 10× accelerated evolution in the first 10 Myr of the simulation. After 3 Myr we selected one simulation in each set in which Enceladus was captured into the 2:1 eE MMR with Dione and acquired a high eccentricity, cloned those simulations, and from that time onward used a much higher tidal response of Enceladus. At 10 Myr we reverted Enceladus to an equilibrium tidal evolution rate and current tidal parameters, evolving the system until 20 Myr or until the Mimas–Tethys resonance reached its present state. We also assumed equilibrium tides on Rhea before 10 Myr and accelerated evolution after than (Section 3.3), and accelerated Titan throughout, but this had little importance as nether Rhea not Titan encountered any resonances.

Table 1. Initial Conditions and Tidal Parameters for the Simulations Shown in Figure 9

MoonEccentricityInclinationRelative TidalTidal k2/Q
Name e i (°)Acceleration×102
Mimas0.0220.110.001
Enceladus0.00350.00310∣10∣10.01∣2.7/3.3∣1
Tethys0.002/0.0030.980.330.02
Dione0.0050.0310.02
Rhea0.0020.331∣1∣50.02
Titan0.03050.33100.3

Note. The slashes separate the values relevant for the two simulations (left-hand side is listed first). For parameters that were changed during the simulation, vertical lines separate values used during 0–3 Myr, 3–10 Myr, and after 10 Myr, respectively. Relative tidal acceleration refers to a factor by which the Saturn's response of k2/Q = (4000)−1 was enhanced at that moon's frequency. Since Q = 100 was typically assumed, the last column is usually equal to the tidal Love number k2, except for Enceladus during the 3–10 Myr period, when k2 = 1 and Q < 100.

Download table as:  ASCIITypeset image

Figure 9 shows the outcomes of two successful simulations (one from each set), which come close to reproducing the current system. Between the two sets of first-stage simulations (20 total), 10 were "successful," meaning that Enceladus and Dione were in their 2:1 eE MMR at 3 Myr. The other 10 simulations were caught in one of the other subresonances of the Enceladus–Dione 2:1 MMR (Figure 8), typically exciting the eccentricities of Tethys or Dione well above the observed values, or "fell out" of the resonance due apparent interaction with secondary resonances. We chose two of the successful simulations for cloning, which was done by sharply changing the tidal properties of Enceladus. The tidal Love number was changed from k2E = 0.01 to k2E = 1, approximating melting. Different clone simulations had the tidal Q of Enceladus in the 30–39 range. This range was chosen through trial and error, as we found that simulations with A = (QS k2E )/(QE k2S ) < 10 led to breaking of the resonance, while those with A > 10 settled into bound cycle within the 2:1 eE subresonance (subscript S designates Saturn's properties at Enceladus's frequency). This critical change in the dynamics of resonance with the change in Enceladus's tidal response was first discovered by Meyer & Wisdom (2008), and our value for critical A agrees with theirs. 4 This phase of the evolution of Enceladus, despite having seemingly chaotic oscillations, is not stochastic but determined by the tidal properties, so all simulations in which Enceladus is dissipative enough (i.e., A > 10) will stay at the threshold of the Enceladus–Dione 2:1 eE resonance indefinitely.

Figure 9.

Figure 9. Two different simulations (plotted on the left- and right-hand sides, respectively) of the recent evolution of the inner moons assuming 10× stronger tides acting on Enceladus in the first 10 Myr of the simulation. Most moons evolved using frequency-independent equilibrium tides with Q/k2 = 4000 for Saturn. Initial conditions are set at about 20 Myr ago, with the goal of reproducing the current system (Table 1). The moons' eccentricities and inclinations were divided into two panels each to show both the high and low values clearly. Vertical dashed lines show the times (3 and 10 Myr) at which the tidal response and/or acceleration of Enceladus were changed (see Table 1). The panels labeled "E-D 2:1" plot the Enceladus–Dione resonant argument 2λD λE ϖE , and panels labeled "M-T 4:2" plot the Mimas–Tethys argument 4λΘ − 2λM − ΩM − ΩΘ.

Standard image High-resolution image

We assume that Enceladus eventually gets out of the resonant mode and migrates due to the "background" tidal response of Saturn that is not dependent on frequency. In reality this transition would be gradual as resonant modes are expected produce a smooth (if narrow) profile of tidal response over frequency (Fuller et al. 2016). We could not simulate this without extensively modifying the integrator and, more importantly, greatly increasing the complexity of our model. At 10 Myr all of our clone simulations switch to Enceladus experiencing Saturn's tides with Q/k2 = 4000 and having its own k2/Q = 0.01, approximately what is expected if the current heating is equal to the observed and is in equilibrium. Out of 20 cloned simulations only three do not complete the transition to a narrow libration within the resonance consistent with observations (in two simulations Enceladus leaves the resonance, in one it is permanently trapped in a large libration state due to secondary resonances). In some simulations the transition is smooth like in the right-hand side of Figure 9, while in the others secondary resonances are encountered during the evolution, leading to temporary increases in the libration amplitude (as in the left-hand side of Figure 9).

The migrations of Mimas, Tethys, and Rhea are largely unaffected by the Enceladus–Dione resonance in our simulations. There is some variation of the eccentricity of Mimas when Enceladus encounters secondary resonances within the 2:1 eE MMR with Dione, but the effect on Mimas is usually minor. The inclination of Mimas is unaffected by any of the dynamics discussed so far, and passes through a series of kicks due to subresonances of the Mimas–Tethys 2:1 commensurability. Subresonances with arguments 2λΘλM + ϖM − 2ΩM , 2λΘλM + ϖM − ΩM − ΩΘ (see Figure 4), and 4λΘ − 2λM − 2ΩM (the pure i2 M term, Figure 3) are encountered at about 3 Myr, 7 Myr, and 14 Myr, respectively. Since the latter two kicks (which are larger) always happen after the simulations were cloned at 3 Myr, we can consider the clones of the original two simulations to be practically independent when it comes to the inclination of Mimas. Out of 18 clones which stayed in the Enceladus–Dione 2:1 MMR, nine were captured into the current Mimas–Tethys 4:2 iM iΘ subresonance (bottom panels in Figure 9). The remaining clones either experienced capture into the iM 2 harmonic or passed the whole forest of subresonances without capture. Therefore we can say that both observed resonances (Mimas–Tethys 4:2 iM iΘ and Enceladus–Dione 2:1 eE ) have high probability (about 50%) when using these initial conditions and assumptions about Enceladus's tidal evolution and internal dissipation. While we can quantify the probabilities of different stochastic outcomes of resonant dynamics, we are currently not able to assess the a priori probability of our initial conditions and assumptions about Saturn's tidal response. Note that our preferred timeline requires Enceladus to settle into the present quiescent state before Mimas acquires a high inclination through its resonance with Dione. The current Mimas–Tethys resonance is very fragile and we often see it broken if it coincides with Enceladus having a very high and/or chaotic eccentricity.

The other parameters of the system at the ends of the simulations in Figure 9 are mostly consistent with the present state. The Mimas–Tethys resonance in every case has a large libration amplitude that is still well short of the observed one (93°), but we expect even very minor subsequent perturbations to modify this quantity (we ignored all moons smaller than Mimas in these runs). The inclinations of Enceladus and Dione are somewhat stochastic but in the range that includes the observed values (iE = 0fdg008 and iD = 0fdg02, respectively), as is the eccentricity of Dione (probably the most variable quantity in our simulations; currently eD = 0.002). For both Mimas and Tethys the inclinations are determined by the initial conditions and their mutual resonance, and the eccentricities depend primarily on the initial values and tidal dissipation, without much stochasticity. However, the eccentricity of Rhea is one quantity that our simulation cannot explain, as we have set it much higher than the observed value in order to avoid the three-body argument of the Enceladus–Dione MMR (Figure 8), but for the tidal dissipation we assume for Rhea cannot subsequently modify that eccentricity. This discrepancy may tell us something about the recent dynamics of Rhea, as we discuss in the next section.

6. Discussion and Conclusions

6.1. Damping of Inclination in Enceladus?

Much of our reasoning that resulted in the model of recent evolution presented in Figure 9 is driven by the survival of the low inclination of Enceladus. In order for Enceladus's inclination to stay so low, both the capture into the 4:2 Enceladus–Dione MMR (Sections 4.1 and 5.1) and crossing of the 11:8 Enceladus–Tethys MMR (Section 4.2) must be avoided. Can we be confident that the inclination of Enceladus was not actually higher in the past and then damped by tides?

The simplest way to estimate the tidal damping of inclination within Enceladus is assuming a homogeneous Enceladus that can be described by a single tidal Love number k2 and a tidal quality factor Q. If we assume that the eccentricity of Enceladus is currently in equilibrium between excitation by resonance with Dione and damping by tides which produce the observed heating (Howett et al. 2011), we get k2/Q ≈ 0.01 (Meyer & Wisdom 2007; Lainey et al. 2012). This corresponds to an eccentricity-damping timescale of about 0.5 Myr, making Enceladus exceptionally dissipative. The timescale for damping of inclination is longer by the factor $7{(\sin i/\sin \theta )}^{2}$, where θ is the forced obliquity of the moon. The forced obliquity of Enceladus has been modeled by Baland et al. (2016) and they found that θ ≤ 4 × 10−4°. This would make $(\sin i/\sin \theta )\geqslant 20$, making the timescale for damping of Enceladus's inclination over 1 Gyr, clearly too slow to affect the dynamics in the time frame we consider.

Another mechanism for inclination damping would be the resonant response of a global ocean (Tyler 2008, 2011). Given that Enceladus likely possesses a subsurface ocean, this possibility must be addressed. Chen et al. (2014) analyzed in depth different mechanisms of tidal dissipation, and found that at Enceladus's present obliquity the heating due to obliquity tides in the ocean is more than three orders of magnitude lower than nonresonant obliquity tides due to the whole-body response of Enceladus. This also implies that the inclination-damping timescale due to resonant ocean tides is longer than 103 Gyr. One important finding by Chen et al. (2014) about resonant tides is that their power depends as a cube of obliquity (as opposed to the square of obliquity in nonresonant tides), meaning that the timescale for inclination damping is inversely proportional to obliquity. However, even if the inclination of Enceladus were about a degree (see Figure 7) and its forced obliquity were θ ≈ 0fdg05, resonant-inclination damping would still be an order of magnitude slower than nonresonant obliquity tides. Finally, even if the parameters used by Chen et al. (2014) were not to apply (for some unforeseen reason) to Enceladus on a high-inclination orbit, their calculations clearly demonstrate that the inclination of Enceladus cannot be brought to its present low value by resonant ocean tides.

The above discussion assumed that Enceladus is in Cassini state 1, in which the forced obliquity due to inclinations is low, and that dissipation within other moons does not affect the evolution of Enceladus's inclination. However, it is in principle possible that the obliquity of a moon could be excited by a spin–orbit resonance. This would be equivalent to the excitation of obliquity of Saturn itself by a secular resonance between the axial precession of Saturn and nodal precession of Neptune's orbit (Hamilton & Ward 2004; Ward & Hamilton 2004). Recently Ćuk et al. (2020) suggested that Uranian moon Oberon may have in the past been caught in a similar resonance with the orbit of Umbriel, and that tidal dissipation within Oberon may have damped the inclination of Umbriel. Using the results of Chen et al. (2014), we surveyed the precession frequencies in the Saturnian system and did not find any candidates for such a resonance, although we cannot exclude this possibility due to uncertainties in the moons' shapes and gravity fields. There are however additional arguments against the relevance of such resonances to Enceladus. Most importantly, the dynamics of the Saturnian satellite system is dominated by Saturn's oblateness, and the mutual perturbations by the moons are less important than in the Uranian system. Therefore there is little or no coupling between the precession frequencies of different moons, and the relatively low mass of Enceladus is particularly unlikely to have a noticeable effect on the rotational dynamics of the larger moons. Furthermore, Ćuk et al. (2020) found that damping of inclination through a spin–orbit resonance can operate only until the relevant orbital inclination damps below the level at which resonance can be maintained, and this level is likely to be higher than the very low current iE = 0fdg008. Therefore we conclude that spin–orbit resonances are very unlikely to have damped the inclination of Enceladus.

6.2. Implications of the Moons' Initial Conditions

Mimas. Our model requires that the eccentricity of Mimas predates the establishment of the current Mimas–Tethys and Enceladus–Dione resonances. The origin of this eccentricity may lie in the past 3:1 Mimas–Dione MMR (see Meyer & Wisdom 2008; Ćuk & El Moutamid 2022), or some other still unidentified, possibly three-body, resonance. The implication is that the damping of Mimas's inclination was limited and therefore that Mimas is unlikely to posses an internal ocean (Rhoden & Walker 2022). A new finding in our paper is that Mimas must have had an ≈0fdg1 inclination before encountering the 2:1 MMR with Tethys. The origin of this inclination is harder to explain, as most three-body resonances do not affect inclination, and a 3:1 MMR crossing with Dione produces a smaller inclination "kick" (Ćuk & El Moutamid 2022). One possibility that needs investigation is whether Dione was more inclined at the time of the Mimas–Dione 3:1 MMR crossing, and that would require producing a self-consistent model of the system's evolution further than 20 Myr into the past.

Enceladus. We find that the eccentricity of Enceladus was already excited (e ≤ 0.005) before encountering the 2:1 MMR with Dione, but its inclination was not. The most likely source of this eccentricity is three-body resonances, either isolated or as part of the 5:3 Dione–Rhea MMR crossing. The apparently pristine inclination of Enceladus implies that it did not encounter any major (first or second order) two-body resonances with mid-sized moons prior to the 2:1 MMR with Dione.

Tethys. Our model requires Tethys to have had a somewhat higher eccentricity (eΘ = 0.002−0.003) 20 Myr ago, and almost its present large inclination. The eccentricity of Tethys poses no challenges, as a past higher eccentricity was suggested on both geophysical (Chen & Nimmo 2008) and dynamical (Ćuk et al. 2016) grounds, and subsequently tidal dissipation is likely to produce Tethys's present low eccentricity. The 1° inclination of Tethys was always known to have to predate the Mimas–Tethys 4:2 resonance (Allan 1969; Sinclair 1972), but its origin was never explained until Ćuk et al. (2016) proposed a Tethys–Dione secular resonance closely following (and dynamically related to) the Dione–Rhea 5:3 MMR crossing. Unless another mechanism of exciting Tethys's inclination is found, our initial conditions effectively require that a passage through the Dione–Rhea 5:3 had happened in the past.

Dione. We start our simulations with Dione that has an excited eccentricity (eD ≈ 0.005) but very low inclination. This is generally consistent with a past secular resonance with Tethys (Ćuk et al. 2016), in which Dione "passed" its inclination and part of its eccentricity (acquired in a resonance with Rhea) to Tethys. The secular resonance is broken when either the eccentricity or inclination of Dione reach very low values (e.g., iD < 0fdg1 for inclination). Therefore, a substantial eccentricity is to be expected to survive if the inclination is very low, implying that the resonance was broken by depletion of Dione's inclination.

Rhea. At the start of our simulation eR = 0.002, 10 times in excess of Rhea's current free eccentricity eR = 2 × 10−4. 5 Rhea's inclination at the start of our simulation is the same as now (iR = 0fdg33) and does not change over the course of it. This substantial inclination is exactly what is expected from a past crossing of the Dione–Rhea 5:3 MMR. This dynamical mechanism is also expected to produce a comparable eccentricity of Rhea, broadly consistent with our initial conditions, but not the observed values. The discrepancy between theoretical expectations and the actual value here is significant, and requires a so-far unknown dynamical mechanism of lowering Rhea's eccentricity. We plan to address this issue in the near future using an integrator that resolves the rotational dynamics of Rhea and search for any additional dynamical features.

6.3. Width and Distribution of Peaks in Tidal Dissipation

Our work is decidedly semiempirical in design, as we acknowledge the importance of the highly variable response of Saturn to the tidal forces of different moons. As we have shown in this paper, both equilibrium tides and resonance-lock-only models that were used so far fail to explain fully the system's dynamics. Apart from the general lack of information on the source of Saturn's dissipation, we tried to keep the number of free parameters to a minimum, leading to our decision to change the evolution rates abruptly. Our decision not to model the interior evolution of Enceladus but to adjust its tidal parameters periodically "by hand" was dictated by our own technical limitations, and we hope that in the future there will be integrated models that fully model both the orbital dynamics and interior evolutions of the moons.

Regardless of the nature and evolution of the resonant modes, it is clear that these are resonant phenomena, and therefore must exhibit a spike-like profile against frequency (possibly Lorentzian or similar). As apparent from Figure 10, we assumed that the width of the normal mode affecting Enceladus to be about 1% of its semimajor axis. We could in principle restrict the very fast evolution to a narrower interval that includes the initial encounter with the Enceladus–Dione 2:1 MMR, but that would require that the resonant mode and the resonance with Dione were encountered at the exact same time due to an unlikely coincidence. Furthermore, if Rhea is only passing though a resonant mode, this mode cannot be too narrow if we were to observe this transient phenomenon. These features in frequency space appear much wider than the resonant modes proposed by Fuller et al. (2016) based on the theory of tidal response in stars and giant planets. We also find that the tidal response at Tethys's frequency must be lower than what appears to be the "background" rate, so this frequency dependence is not restricted to peaks in dissipation, but also has local minima ("troughs"). Apart from being relatively wide, resonant modes cannot be too few and far in between if they were recently encountered by both Enceladus and Rhea as we propose.

Figure 10.

Figure 10. Semimajor axis of Enceladus (solid green line) in the right-hand simulation in Figure 9, compared to locations of MMRs with other moons. Solid lines are from this numerical simulation, while dashed lines are analytical estimates. The blue dashed line shows the location of the 1:2 resonance with Janus or its progenitor, assuming orbital evolution due ring torques (Tajeddine et al. 2017). The solid red line is the location of the 11:8 Enceladus–Tethys resonance in our simulation. The purple dashed line extrapolates the location of the 2:1 resonance with Dione into the past assuming equilibrium tides only. The green dashed line extrapolates the semimajor axis of Enceladus before the simulation starts assuming equilibrium tides.

Standard image High-resolution image

The determination of a consensus result for the current rate of tidal evolution of Titan is of great importance, as it will give us a major indication if resonance locking is present or not. Of all the moons of Saturn that produce significant tides (Mimas–Titan), Titan is most likely to be in a resonant lock, as it has the slowest (equilibrium) tidal evolution and is the least likely to have been recently reaccreted in some kind of late cataclysm. If Titan is not in a resonance lock, it is possible that the resonant modes are moving inward (in terms of semimajor axis at which they are encountered), making it somewhat less surprising that encounters between moons and modes are common. Still, two moons encountering resonant modes within the last few tens of millions of years appears unlikely unless the modes are very numerous, or their distributions are correlated with those of the moons. The last possibility may indicate that the moons may have reaccreted close to the modes, either due to where the last generation of satellites was before the assumed instability, or for some other reason. The only certainty is that the tidal evolution of the Saturnian system holds yet more surprises for us.

6.4. Summary

In this work we tried to consider all of the available constraints on the recent (last few tens of millions of years) and current orbital evolution of the major Saturnian satellites. We find that no single mechanism of orbital evolution proposed so far (including frequency-independent equilibrium tides and the evolution through resonance locking) can explain the orbits of these moons. Strong equilibrium tides can explain the existence of the observed resonances (Mimas–Tethys and Enceladus–Dione) and the current heating rate of Enceladus, while resonant modes are necessary to explain the current dynamics of Rhea and the original capture of Enceladus into the resonance with Dione. Additionally, the evolution of Tethys needs to be slower than that of other moons, implying "troughs" as well as "peaks" in the response as a function of frequency.

In order to reproduce the encounter of Enceladus with the 2:1 resonance with Dione successfully we require a passage through a resonant mode, rather than locking to a resonant mode. This is reasonable if the resonant modes in the inner system are evolving more slowly than those at larger distances, as originally predicted Fuller et al. (2016), but would not work in the context of inertial waves (Lainey et al. 2020). Alternatively, if Titan is currently not locked to a resonant mode as the results of Jacobson (2022) suggest, it is also possible that the resonant modes move inward, and in that case the moons could only temporarily cross the modes, rather than become locked to them.

Given the complexity and uncertainties of the tidal evolution rates that not only vary from moon to moon but also over time, it is difficult to reach firm conclusions about the age of the system. However, given the amount of dynamical excitation that the inner moons (especially Mimas and Enceladus) may have experienced in the last 20 Myr, it is difficult to envision that this system of relatively "dynamically cold" satellites evolved this way for hundreds of millions of years, let alone multiple gigayears. We hope that more precise future determinations of the current orbital evolution rates of the Saturnian moons (based on astrometry or spacecraft data) will be able to confirm or falsify our model of their recent evolution.

Acknowledgments

This work was supported by NASA Solar System Workings Program awards 80NSSC19K0544 (to M.Ć. and M.E.M.) and 80NSSC22K0979 (to M.Ć.). We would like to thank Jim Fuller, Valery Lainey, Bob Jacobson, and Francis Nimmo for very insightful discussions. We also thank the International Space Science Institute in Bern for organizing an extremely useful workshop in the evolution of the Saturnian system (2022 May). We wish to thank the two anonymous reviewers whose comments greatly improved the paper.

Footnotes

  • 3  

    We assumed a value on the high-end of this estimate of dissipation in order to make the simulations faster.

  • 4  

    As Meyer & Wisdom (2008) explored a more standard case in which the orbital evolution of Dione is not negligible compared to that of Enceladus, A in their paper must be multiplied by ${(1-({\dot{a}}_{D}{a}_{E})/({\dot{a}}_{E}{a}_{D}))}^{-1}$.

  • 5  

    The total eccentricity of Rhea is larger due to a dominant term forced by Titan which is not relevant for the dynamics discussed here.

Please wait… references are loading.
10.3847/PSJ/acde80