Nearby Supernova and Cloud Crossing Effects on the Orbits of Small Bodies in the Solar System

Leeanne Smith Department of Atmospheric Sciences, University of Illinois, Urbana, IL 61801, USA Jesse A. Miller Department of Astronomy, Boston University, Boston, MA 02215, USA Department of Astronomy, University of Illinois, Urbana, IL 61801, USA Brian D. Fields Department of Astronomy, University of Illinois, Urbana, IL 61801, USA
Abstract

Supernova blasts envelop many surrounding stellar systems, transferring kinetic energy to small bodies in the systems. Geologic evidence from Fe60superscriptFe60{}^{60}{\rm Fe}start_FLOATSUPERSCRIPT 60 end_FLOATSUPERSCRIPT roman_Fe points to recent nearby supernova activity within the past several Myr. Here, we model the transfer of energy and resulting orbital changes from these supernova blasts to the Oort Cloud, the Kuiper belt, and Saturn’s Phoebe ring. For the Oort Cloud, an impulse approximation shows that a 50 pc supernova can eject approximately half of all objects less than 1 cm while altering the trajectories of larger ones, depending on their orbital parameters. For stars closest to supernovae, objects up to 100similar-toabsent100\sim 100∼ 100 m can be ejected. Turning to the explored solar system, we find that supernovae closer than 50 pc may affect Saturn’s Phoebe ring and can sweep away Kuiper belt dust. It is also possible that the passage of the solar system through a dense interstellar cloud could have a similar effect; a numerical trajectory simulation shows that the location of the dust grains and the direction of the wind (from a supernova or interstellar cloud) has a significant impact on whether or not the grains will become unbound from their orbit in the Kuiper belt. Overall, nearby supernovae sweep micron-sized dust from the solar system, though whether the grains are ultimately cast towards the Sun or altogether ejected depends on various factors. Evidence of supernova-modified dust grain trajectories may be observed by New Horizons, though further modeling efforts are required.

Supernovae (1668); Kuiper Belt (893); Oort Cloud (1157); Saturn (1426)
software: astropy (Robitaille et al., 2013; Price-Whelan et al., 2018), numpy (Harris et al., 2020), matplotlib (Hunter, 2007), scipy (Virtanen et al., 2020)

1 Introduction

Measurements of live Fe60superscriptFe60{}^{60}{\rm Fe}start_FLOATSUPERSCRIPT 60 end_FLOATSUPERSCRIPT roman_Fe (t1/2=2.6subscript𝑡122.6t_{1/2}=2.6italic_t start_POSTSUBSCRIPT 1 / 2 end_POSTSUBSCRIPT = 2.6 Myr) in geological deposits show that Earth was in the vicinity of supernovae (SNe) 3similar-toabsent3\sim 3∼ 3 Myr and similar-to\sim 7 Myr ago (Knie et al., 1999, 2004; Fitoussi et al., 2008; Ludwig et al., 2016; Wallner et al., 2016, 2021). Freshly synthesized Fe60superscriptFe60{}^{60}{\rm Fe}start_FLOATSUPERSCRIPT 60 end_FLOATSUPERSCRIPT roman_Fe from the explosion was incorporated into dust grains, which propagated through interstellar space and the solar system, eventually landing on Earth (Athanassiadou & Fields, 2011; Fry et al., 2020).

While many studies have investigated the effects of supernovae on the Earth (e.g., Ellis et al., 1996; Gehrels et al., 2003; Thomas et al., 2016), there have been far fewer investigations into effects throughout the solar system. Stern & Shull (1988) and Stern (1990) calculated how SNe could flash heat the surface of Oort Cloud bodies and how dynamical drag could erode their surfaces. Armed with the evidence of recent supernova activity, we are motivated to consider a wider variety of ways SNe may have affected our solar system. In this work, we consider the possibility that nearby SNe may have dynamically altered the orbits of small solar system bodies.

The Oort Cloud, being distant from the Sun and therefore only tenuously gravitationally bound (Oort, 1950), keenly experiences extrasolar influences. These influences include the Galactic tidal field (Heisler & Tremaine, 1986) and nearby passing stars (Rickman et al., 2008). In addition, the gravitational perturbation of passing giant interstellar clouds has been investigated (Jakubík & Neslušan, 2008; Hut & Tremaine, 1985), though not the direct gasdynamical impulse of these clouds.

Recently, Opher et al. (2024) proposed that the solar system travelled through a dense cold cloud called the Local Lynx of Cold Cloud (LxCC) near the same time as the initial onset of the Fe60superscriptFe60{}^{60}{\rm Fe}start_FLOATSUPERSCRIPT 60 end_FLOATSUPERSCRIPT roman_Fe pulse 3 Myr ago. The dense cloud (nH3000similar-tosubscript𝑛H3000n_{\rm H}\sim 3000italic_n start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT ∼ 3000 cm-3) would have compressed the heliosphere to a mere 0.22 au, stripping the entire solar system of its protective solar wind.

Both the direct SN blast wave and cold cloud scenarios call for a significant shift in ram pressure felt throughout the solar system. While a SN blast does not change the density of the local interstellar medium (ISM) greatly (at most a factor of 4 due to a strong shock), the speed increases by a factor of similar-to\sim10, depending on the distance to the SN. In contrast, the cold cloud enhances ISM density by 4-5 orders of magnitude with little change in velocity.

The SN blast wave and cold clouds scenarios also differ in the regions they influence. The cold cloud coats the entire solar system, even Mercury’s orbit (Opher et al., 2024). The effects of the SN blast, on the other hand, are highly dependent on the distance to the SN. Estimates for the distance to the 3 Myr-old SN vary depending on many factors such as the total Fe60superscriptFe60{}^{60}{\rm Fe}start_FLOATSUPERSCRIPT 60 end_FLOATSUPERSCRIPT roman_Fe abundance generated by the SN, amount of Fe60superscriptFe60{}^{60}{\rm Fe}start_FLOATSUPERSCRIPT 60 end_FLOATSUPERSCRIPT roman_Fe incorporated into dust grains, and Fe60superscriptFe60{}^{60}{\rm Fe}start_FLOATSUPERSCRIPT 60 end_FLOATSUPERSCRIPT roman_Fe uptake into geological samples (Ertel et al., 2023). Based on Fe60superscriptFe60{}^{60}{\rm Fe}start_FLOATSUPERSCRIPT 60 end_FLOATSUPERSCRIPT roman_Fe measurements, the SN distance was approximately 201402014020-14020 - 140 pc from Earth (Ertel & Fields, 2023; Fry et al., 2015). Miller & Fields (2022) performed simulations of how the heliosphere responds to SNe at different distances, finding that a SN at a moderate distance of 50 pc compresses the heliosphere to 10 au, the orbit of Saturn. Therefore, the effects of a SN blast wave are not felt as widely throughout the solar system as those from a cold cloud. Observable traces of the SN blast within the solar system could potentially be used as a way to indicate SN distance, or even distinguish which of these two cases occurred.

To this end, we investigate the effects of how a large ram pressure from the SN blast or cold cloud could influence solar system orbits. This examination is done in three cases: section 2 describes an impulse approximation for Oort Cloud bodies, and section 3 considers the effects on Saturn’s rings and Kuiper belt dust. Finally, discussion and conclusions are given in section 4.

2 Oort Cloud

2.1 Impulse Approximation

Due to the significant difference between Oort Cloud orbital time (1 Myr at 104superscript10410^{4}10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT au) and duration of the SN blast in the Sedov solution (similar-to\sim0.05 Myr), the SN blast wave imparts an impulsive force onto the Oort Cloud. Rather than assuming a time-dependent force, it is valid to simply assume that the blast wave imparts an instantaneous velocity boost to the comets. Assuming the comet absorbs all incident kinetic energy from the blast, the comet gains a kick velocity of

vkick=38πESNρcrcD2,subscript𝑣kick38𝜋subscript𝐸𝑆𝑁subscript𝜌𝑐subscript𝑟𝑐superscript𝐷2v_{\rm kick}=\sqrt{{\frac{3}{8\pi}}\frac{E_{SN}}{\rho_{c}r_{c}D^{2}}},italic_v start_POSTSUBSCRIPT roman_kick end_POSTSUBSCRIPT = square-root start_ARG divide start_ARG 3 end_ARG start_ARG 8 italic_π end_ARG divide start_ARG italic_E start_POSTSUBSCRIPT italic_S italic_N end_POSTSUBSCRIPT end_ARG start_ARG italic_ρ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT italic_D start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG , (1)

where ESNsubscript𝐸𝑆𝑁E_{SN}italic_E start_POSTSUBSCRIPT italic_S italic_N end_POSTSUBSCRIPT is the energy of the supernova (assumed to be 1051superscript105110^{51}10 start_POSTSUPERSCRIPT 51 end_POSTSUPERSCRIPT erg), rcsubscript𝑟𝑐r_{c}italic_r start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT and ρcsubscript𝜌𝑐\rho_{c}italic_ρ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT are the radius and density of the comet, and D𝐷Ditalic_D is the distance from the supernova.

This is in no way intended to be an exact description of the modern distribution of Oort Cloud comets, as such information is outside the scope of this paper. Rather, our purpose is to describe what would happen to such comets when they are hit by a SN blast.

These simulations consist of 106superscript10610^{6}10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT Oort Cloud objects with radii ranging from 10μm – 10010𝜇m – 10010\ \mu\textrm{m -- }10010 italic_μ m – 100 m and densities ranging from 0.51.5 g cm30.51.5superscript g cm30.5\textrm{--}1.5\textrm{ g cm}^{-3}0.5 – 1.5 g cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT. The orbital parameters, as well as size and density, are all chosen independently. This wide radius range allows us to capture the dynamics of all objects from dust grains to modest asteroids. Consistent with observed comets, they are given highly eccentric orbits (0.9<e<10.9𝑒10.9<e<10.9 < italic_e < 1) with semi-major axes within 10,000<a<20,000formulae-sequence10000𝑎2000010,000<a<20,00010 , 000 < italic_a < 20 , 000 au (Dones et al., 2015). Each comet has randomly generated Keplerian coordinates, including the angle of inclination i𝑖iitalic_i, argument of periapsis ω𝜔\omegaitalic_ω, right ascension of the ascending node ΩΩ\Omegaroman_Ω, and mean anomaly ν𝜈\nuitalic_ν within ranges of [0,π)0𝜋[0,\pi)[ 0 , italic_π ), [0,2π)02𝜋[0,2\pi)[ 0 , 2 italic_π ), [0,2π)02𝜋[0,2\pi)[ 0 , 2 italic_π ), and [0,2π)02𝜋[0,2\pi)[ 0 , 2 italic_π ) respectively.

Each comet was converted to Cartesian coordinates, given an additional vkicksubscript𝑣𝑘𝑖𝑐𝑘v_{kick}italic_v start_POSTSUBSCRIPT italic_k italic_i italic_c italic_k end_POSTSUBSCRIPT in the vxsubscript𝑣𝑥v_{x}italic_v start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT direction, and converted back to Keplerian coordinates to analyze orbital properties. Note that the direction of the kick itself is unimportant, as the spherical nature of the Oort Cloud ensures a symmetric geometry, so results will be independent of direction.

Due to the uncertainty in the distance to the SN, we employ a set of calculations for four SNe at distances of 10, 20, 50, and 100 pc.

2.2 Results

The initial distribution and results of applying the kick velocity of a SN 10 pc distant are shown in Fig. 1. For the semi-major axis, there are significantly fewer objects within the initial semi-major axis range because the rest were either pushed to a different orbit or completely ejected (which is treated as a negative semi-major axis). As will be shown later, the amount the semi-major axis changes is mostly determined by the body’s mass and only weakly dependent on the orbital parameters. The eccentricity reveals that while many bodies became more circularized, many others were unbound with e1much-greater-than𝑒1e\gg 1italic_e ≫ 1 and a<0𝑎0a<0italic_a < 0.

Refer to caption
Figure 1: Histograms of orbital parameters of Oort Cloud comets in their initial distribution (blue) and after being hit by a remarkably close supernova 10 pc away (orange). The parameters shown are (left to right, top to bottom) semi-major axis, eccentricity, angle of inclination, right ascension of the ascending node, argument of periapsis, and the true anomaly.

The three angular coordinates (i𝑖iitalic_i, ΩΩ\Omegaroman_Ω, and ω𝜔\omegaitalic_ω) have more interesting features. In these plots, unlike those of semi-major axis and eccentricity, all orbits are represented. The inclination shows a slight preference to move towards i=0𝑖0i=0italic_i = 0 and π𝜋\piitalic_π. There is a strong preference to shift the right ascension towards Ω=0Ω0\Omega=0roman_Ω = 0, π𝜋\piitalic_π, and 2π2𝜋2\pi2 italic_π and the argument of periapsis towards ω=π/2𝜔𝜋2\omega=\pi/2italic_ω = italic_π / 2 and 3π/23𝜋23\pi/23 italic_π / 2. The large spikes are the result of becoming unbound: unbound bodies are the closest they get to the reference plane that defines i=0𝑖0i=0italic_i = 0, and so the point at which they cross this plane (the right ascension) is near their current position. The argument of periapsis is strongly peaked because the unbound bodies are very near their periapsis and travelling in a near-straight line, which corresponds to being at the stated values of ω𝜔\omegaitalic_ω. Lastly, the true anomaly is clustered towards 0 for similar reasons: unbound particles are very near their periapsis, which is defined as the point at which ν=0𝜈0\nu=0italic_ν = 0.

The simulated objects were sorted and grouped into logarithmically-spaced radius bins. For each bin, the percentage of orbits that became unbound was calculated. The result is shown in Figure 2. As expected, the unbound fraction shows a strong dependence on size, with the lighter particles being completely ejected while larger ones barely move. This is also highly dependent on SN distance. For SNe at 10, 20, 50, and 100 pc away, we see steep drop offs in the number of unbound objects with radii around 20, 5, 0.8, and 0.2 cm respectively.

We also compare our results to those of Stern (1990), who performs a similar calculation, albeit without the complete description of 3D orbital mechanics utilized here. That study finds that all particles at a>104𝑎superscript104a>10^{4}italic_a > 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT au with rc<subscript𝑟𝑐absentr_{c}<italic_r start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT < 0.3 mm are ejected by a SN at 40 pc. We find a fairly good agreement for the particles that become completely unbound. However, there are a large number of particles that can become unbound due to their orbital parameters that Stern (1990) does not account for.

We see that at distances appropriate for the recent supernovae (50100similar-toabsent50100\sim 50-100∼ 50 - 100 pc), Oort Cloud particles of 1cmsimilar-toabsent1cm\sim 1\ \rm cm∼ 1 roman_cm will be strongly affected, with some ejected to interstellar space. Other such particles will remain bound, in orbits coming much closer to the Sun; we have calculated the perihelion distribution of the perturbed particles and find that only a tiny fraction of the particles reach within 100 au. For supernovae close enough to be dangerous (10-20 pc), objects as large as 1msimilar-toabsent1m\sim 1\ \rm m∼ 1 roman_m will suffer these effects.

Refer to caption
Figure 2: Fraction of comets of a given size that become unbound from a SN blast, evaluated for SN distances of 10, 20, 50, and 100 pc from the solar system. The dashed line comes from a simplified calculation by Stern (1990) for a SN 40 pc away for particles orbiting 20,000 au from the Sun.

3 Saturn’s Phoebe Ring and Kuiper Belt Dust

Unlike in the Oort Cloud, where the 0.05similar-toabsent0.05\sim 0.05∼ 0.05 Myr passage of a SN blast could be modelled as an impulse approximation, bodies that make up planetary rings and the Kuiper belt orbit much more rapidly. The total force imparted by the blast spans many orbits rather than a single instant. To investigate these orbital changes, a more complete model is required. Therefore, we make estimates using a secular approximation, followed by a numerical implementation.

Two cases stand out as particularly enlightening: Saturn’s rings and Kuiper belt dust. Saturn’s A and B rings have a short dynamical timescale and high particle densities, making collisions frequent. The collision timescale is much shorter than the timescale for the wind perturbations to become effective, so that these rings will be unaffected by the SN.111We thank the anonymous referee for pointing this out. However, Saturn’s largest ring, the Phoebe ring, has a much lower number density, and grain sizes are approximately 10 μ𝜇\muitalic_μm (Verbiscer et al., 2009), making the Phoebe ring a valid target for our calculation. Similarly, Kuiper belt dust represents tiny bodies that keenly experience an outside force, and their modified trajectories can be compared to spacecraft measurements today (Poppe et al., 2019).

3.1 Secular Approximation: Formalism

The trajectories of small grains in the solar system are influenced by a drag force,

Fdrag(t)=Cdρwπrgr2|vrel|2v^relsubscript𝐹drag𝑡subscript𝐶dsubscript𝜌w𝜋superscriptsubscript𝑟gr2superscriptsubscript𝑣rel2subscript^𝑣relF_{\rm drag}(t)=C_{\rm d}\ \rho_{\rm w}\ \pi r_{\rm gr}^{2}\ |\vec{v}_{\rm rel% }|^{2}\ \hat{v}_{\rm rel}italic_F start_POSTSUBSCRIPT roman_drag end_POSTSUBSCRIPT ( italic_t ) = italic_C start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT roman_w end_POSTSUBSCRIPT italic_π italic_r start_POSTSUBSCRIPT roman_gr end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT | over→ start_ARG italic_v end_ARG start_POSTSUBSCRIPT roman_rel end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT over^ start_ARG italic_v end_ARG start_POSTSUBSCRIPT roman_rel end_POSTSUBSCRIPT (2)

Here, vrel=vwusubscript𝑣𝑟𝑒𝑙subscript𝑣w𝑢\vec{v_{rel}}=\vec{v}_{\rm w}-\vec{u}over→ start_ARG italic_v start_POSTSUBSCRIPT italic_r italic_e italic_l end_POSTSUBSCRIPT end_ARG = over→ start_ARG italic_v end_ARG start_POSTSUBSCRIPT roman_w end_POSTSUBSCRIPT - over→ start_ARG italic_u end_ARG gives the difference between the wind velocity vwsubscript𝑣w\vec{v}_{\rm w}over→ start_ARG italic_v end_ARG start_POSTSUBSCRIPT roman_w end_POSTSUBSCRIPT and the particle velocity u𝑢\vec{u}over→ start_ARG italic_u end_ARG; this relative velocity has magnitude |vrel|subscript𝑣rel|\vec{v}_{\rm rel}|| over→ start_ARG italic_v end_ARG start_POSTSUBSCRIPT roman_rel end_POSTSUBSCRIPT | and direction v^relsubscript^𝑣rel\hat{v}_{\rm rel}over^ start_ARG italic_v end_ARG start_POSTSUBSCRIPT roman_rel end_POSTSUBSCRIPT. We take the drag coefficient Cd=1subscript𝐶d1C_{\rm d}=1italic_C start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT = 1. Finally, ρwsubscript𝜌w\rho_{\rm w}italic_ρ start_POSTSUBSCRIPT roman_w end_POSTSUBSCRIPT is the wind density and rgrsubscript𝑟grr_{\rm gr}italic_r start_POSTSUBSCRIPT roman_gr end_POSTSUBSCRIPT is the grain radius.

This leads to a slow (secular) change of the orbital parameters for each particle. Pástor et al. (2011) derived the expression governing these changes, time-averaged and to leading order in the perturbations. We focus on the case when vw/u1much-greater-thansubscript𝑣w𝑢1v_{\rm w}/u\gg 1italic_v start_POSTSUBSCRIPT roman_w end_POSTSUBSCRIPT / italic_u ≫ 1, which is valid for the high velocity of a SN blast. The semi-major axis evolves as

dadt=2Cdnwmwmgrπrgr2vwaf𝑑𝑎𝑑𝑡2subscript𝐶dsubscript𝑛wsubscript𝑚wsubscript𝑚gr𝜋superscriptsubscript𝑟gr2subscript𝑣w𝑎𝑓\frac{da}{dt}=-2\ C_{\rm d}\ n_{\rm w}\frac{m_{\rm w}}{m_{\rm gr}}\pi r_{\rm gr% }^{2}v_{\rm w}\ a\ fdivide start_ARG italic_d italic_a end_ARG start_ARG italic_d italic_t end_ARG = - 2 italic_C start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT roman_w end_POSTSUBSCRIPT divide start_ARG italic_m start_POSTSUBSCRIPT roman_w end_POSTSUBSCRIPT end_ARG start_ARG italic_m start_POSTSUBSCRIPT roman_gr end_POSTSUBSCRIPT end_ARG italic_π italic_r start_POSTSUBSCRIPT roman_gr end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_v start_POSTSUBSCRIPT roman_w end_POSTSUBSCRIPT italic_a italic_f (3)

where we have separated the wind density into ρw=mwnwsubscript𝜌wsubscript𝑚wsubscript𝑛w\rho_{\rm w}=m_{\rm w}n_{\rm w}italic_ρ start_POSTSUBSCRIPT roman_w end_POSTSUBSCRIPT = italic_m start_POSTSUBSCRIPT roman_w end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT roman_w end_POSTSUBSCRIPT, and mgrsubscript𝑚grm_{\rm gr}italic_m start_POSTSUBSCRIPT roman_gr end_POSTSUBSCRIPT is the grain mass. f𝑓fitalic_f is a dimensionless function that depends on the eccentricity and angular orbital parameters, and the orientation of the wind relative to the orbital plane; in our estimate we will take it to be f1similar-to𝑓1f\sim 1italic_f ∼ 1. We see from eq. (3) that the characteristic timescale for the change in the semi-major axis is

Γa=|da/dt|a 2Cdnwmwmgrπrgr2FdragmgrvwsubscriptΓ𝑎𝑑𝑎𝑑𝑡𝑎similar-to2subscript𝐶dsubscript𝑛wsubscript𝑚wsubscript𝑚gr𝜋superscriptsubscript𝑟gr2subscript𝐹dragsubscript𝑚grsubscript𝑣w\Gamma_{a}=\frac{|da/dt|}{a}\ \sim\ 2\ C_{\rm d}\ n_{\rm w}\frac{m_{\rm w}}{m_% {\rm gr}}\pi r_{\rm gr}^{2}\frac{F_{\rm drag}}{m_{\rm gr}v_{\rm w}}roman_Γ start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT = divide start_ARG | italic_d italic_a / italic_d italic_t | end_ARG start_ARG italic_a end_ARG ∼ 2 italic_C start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT roman_w end_POSTSUBSCRIPT divide start_ARG italic_m start_POSTSUBSCRIPT roman_w end_POSTSUBSCRIPT end_ARG start_ARG italic_m start_POSTSUBSCRIPT roman_gr end_POSTSUBSCRIPT end_ARG italic_π italic_r start_POSTSUBSCRIPT roman_gr end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT divide start_ARG italic_F start_POSTSUBSCRIPT roman_drag end_POSTSUBSCRIPT end_ARG start_ARG italic_m start_POSTSUBSCRIPT roman_gr end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT roman_w end_POSTSUBSCRIPT end_ARG (4)

which we see is just twice the ratio of the drag force on a grain divided by the grain momentum mgrvwsubscript𝑚grsubscript𝑣wm_{\rm gr}v_{\rm w}italic_m start_POSTSUBSCRIPT roman_gr end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT roman_w end_POSTSUBSCRIPT as seen by the wind. This gives a characteristic decay timescale

τasubscript𝜏𝑎\displaystyle\tau_{a}italic_τ start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT =\displaystyle== 1Γa1subscriptΓ𝑎\displaystyle\frac{1}{\Gamma_{a}}divide start_ARG 1 end_ARG start_ARG roman_Γ start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT end_ARG (5)
=\displaystyle== 5Myr(500km/svw)(102cm3nw)(1μmrgr)25Myr500kmssubscript𝑣wsuperscript102superscriptcm3subscript𝑛wsuperscript1𝜇msubscript𝑟gr2\displaystyle 5\ {\rm Myr}\left(\frac{500\ {\rm km/s}}{v_{\rm w}}\right)\left(% \frac{10^{-2}\ {\rm cm}^{-3}}{n_{\rm w}}\right)\left(\frac{1\ \mu{\rm m}}{r_{% \rm gr}}\right)^{2}5 roman_Myr ( divide start_ARG 500 roman_km / roman_s end_ARG start_ARG italic_v start_POSTSUBSCRIPT roman_w end_POSTSUBSCRIPT end_ARG ) ( divide start_ARG 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT end_ARG start_ARG italic_n start_POSTSUBSCRIPT roman_w end_POSTSUBSCRIPT end_ARG ) ( divide start_ARG 1 italic_μ roman_m end_ARG start_ARG italic_r start_POSTSUBSCRIPT roman_gr end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (6)

for a grain of density ρgr=2g/cm3subscript𝜌gr2gsuperscriptcm3\rho_{\rm gr}=2\ \rm g/cm^{3}italic_ρ start_POSTSUBSCRIPT roman_gr end_POSTSUBSCRIPT = 2 roman_g / roman_cm start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT, which is much longer than the 0.05Myrsimilar-toabsent0.05Myr\sim 0.05\ \rm Myr∼ 0.05 roman_Myr timescale a SN blast will extend to the location of Saturn or the Kuiper belt (Miller & Fields, 2022).

For the grain orbital eccentricity, Pástor et al. (2011) find the time-averaged change to be

dedt𝑑𝑒𝑑𝑡\displaystyle\frac{de}{dt}divide start_ARG italic_d italic_e end_ARG start_ARG italic_d italic_t end_ARG =\displaystyle== Cdnwmwmgrvw(g1+vwvcg2)subscript𝐶dsubscript𝑛wsubscript𝑚wsubscript𝑚grsubscript𝑣wsubscript𝑔1subscript𝑣wsubscript𝑣𝑐subscript𝑔2\displaystyle-C_{\rm d}n_{\rm w}\frac{m_{\rm w}}{m_{\rm gr}}v_{\rm w}\left(g_{% 1}+\frac{v_{\rm w}}{v_{c}}g_{2}\right)- italic_C start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT roman_w end_POSTSUBSCRIPT divide start_ARG italic_m start_POSTSUBSCRIPT roman_w end_POSTSUBSCRIPT end_ARG start_ARG italic_m start_POSTSUBSCRIPT roman_gr end_POSTSUBSCRIPT end_ARG italic_v start_POSTSUBSCRIPT roman_w end_POSTSUBSCRIPT ( italic_g start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + divide start_ARG italic_v start_POSTSUBSCRIPT roman_w end_POSTSUBSCRIPT end_ARG start_ARG italic_v start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG italic_g start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) (7)
=\displaystyle== 12Γa(g1+vwvcg2)12subscriptΓ𝑎subscript𝑔1subscript𝑣wsubscript𝑣𝑐subscript𝑔2\displaystyle-\frac{1}{2}\Gamma_{a}\left(g_{1}+\frac{v_{\rm w}}{v_{c}}g_{2}\right)- divide start_ARG 1 end_ARG start_ARG 2 end_ARG roman_Γ start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ( italic_g start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + divide start_ARG italic_v start_POSTSUBSCRIPT roman_w end_POSTSUBSCRIPT end_ARG start_ARG italic_v start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG italic_g start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) (8)

where the dimensionless factors g1subscript𝑔1g_{1}italic_g start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and g2subscript𝑔2g_{2}italic_g start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT are functions of the eccentricity, angular orbital parameters, and wind orientation. We see that there are two terms, the first of which has the same order of magnitude as ΓasubscriptΓ𝑎\Gamma_{a}roman_Γ start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT, and thus will change over the same timescale. The second term differs by the ratio of the wind speed to the circular orbit speed vc=GM/asubscript𝑣𝑐𝐺𝑀𝑎v_{c}=\sqrt{GM/a}italic_v start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = square-root start_ARG italic_G italic_M / italic_a end_ARG. We see that the eccentricity decay rate is of order

Γe12vwvcΓasimilar-tosubscriptΓ𝑒12subscript𝑣wsubscript𝑣𝑐subscriptΓ𝑎\Gamma_{e}\sim\frac{1}{2}\frac{v_{\rm w}}{v_{c}}\Gamma_{a}roman_Γ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ∼ divide start_ARG 1 end_ARG start_ARG 2 end_ARG divide start_ARG italic_v start_POSTSUBSCRIPT roman_w end_POSTSUBSCRIPT end_ARG start_ARG italic_v start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG roman_Γ start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT (9)

and thus the timescale for eccentricity change due to this term is

τe=1|de/dt|2vcvwτasubscript𝜏𝑒1𝑑𝑒𝑑𝑡similar-to2subscript𝑣𝑐subscript𝑣wsubscript𝜏𝑎\tau_{e}=\frac{1}{|de/dt|}\sim 2\frac{v_{c}}{v_{\rm w}}\tau_{a}italic_τ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG | italic_d italic_e / italic_d italic_t | end_ARG ∼ 2 divide start_ARG italic_v start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG start_ARG italic_v start_POSTSUBSCRIPT roman_w end_POSTSUBSCRIPT end_ARG italic_τ start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT (10)

and finally,

τeτasubscript𝜏𝑒subscript𝜏𝑎\displaystyle\frac{\tau_{e}}{\tau_{a}}divide start_ARG italic_τ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT end_ARG start_ARG italic_τ start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT end_ARG =\displaystyle== 2vcvw2subscript𝑣𝑐subscript𝑣w\displaystyle 2\frac{v_{c}}{v_{\rm w}}2 divide start_ARG italic_v start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG start_ARG italic_v start_POSTSUBSCRIPT roman_w end_POSTSUBSCRIPT end_ARG (11)
similar-to\displaystyle\sim 0.02(500km/svw)(MM)1/2(40aua)1/20.02500kmssubscript𝑣wsuperscript𝑀subscript𝑀direct-product12superscript40au𝑎12\displaystyle 0.02\ \left(\frac{500\ {\rm km/s}}{v_{\rm w}}\right)\left(\frac{% M}{M_{\odot}}\right)^{1/2}\left(\frac{40\ {\rm au}}{a}\right)^{1/2}0.02 ( divide start_ARG 500 roman_km / roman_s end_ARG start_ARG italic_v start_POSTSUBSCRIPT roman_w end_POSTSUBSCRIPT end_ARG ) ( divide start_ARG italic_M end_ARG start_ARG italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT ( divide start_ARG 40 roman_au end_ARG start_ARG italic_a end_ARG ) start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT (12)

and we see that for SN conditions, the eccentricity will change more rapidly than the semi-major axis (eq. 5). For the semi-major axis in eq. 5, this amounts to τe0.1similar-tosubscript𝜏𝑒0.1\tau_{e}\sim 0.1italic_τ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ∼ 0.1 Myr.

The angle of inclination i𝑖iitalic_i has a similar form as the eccentricity and so may be treated in a similar manner. This rate of change and corresponding timescale, then, are

didt𝑑𝑖𝑑𝑡\displaystyle\frac{di}{dt}divide start_ARG italic_d italic_i end_ARG start_ARG italic_d italic_t end_ARG =\displaystyle== 14Γa(h1+vwvch2)14subscriptΓ𝑎subscript1subscript𝑣wsubscript𝑣𝑐subscript2\displaystyle-\frac{1}{4}\Gamma_{a}\left(h_{1}+\frac{v_{\rm w}}{v_{c}}h_{2}\right)- divide start_ARG 1 end_ARG start_ARG 4 end_ARG roman_Γ start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ( italic_h start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + divide start_ARG italic_v start_POSTSUBSCRIPT roman_w end_POSTSUBSCRIPT end_ARG start_ARG italic_v start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG italic_h start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) (13)
τisubscript𝜏𝑖\displaystyle\tau_{i}italic_τ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT similar-to\displaystyle\sim 4vcvwτa4subscript𝑣𝑐subscript𝑣wsubscript𝜏𝑎\displaystyle 4\frac{v_{c}}{v_{\rm w}}\tau_{a}4 divide start_ARG italic_v start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG start_ARG italic_v start_POSTSUBSCRIPT roman_w end_POSTSUBSCRIPT end_ARG italic_τ start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT (14)

where h1subscript1h_{1}italic_h start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and h2subscript2h_{2}italic_h start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT are defined in the same manner as g1subscript𝑔1g_{1}italic_g start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and g2subscript𝑔2g_{2}italic_g start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT. The inclination timescale is merely a factor of 2 larger than the eccentricity timescale.

3.2 Secular Approximation: Results

Equations (5), (10), and (14) can be used to derive the timescales on which orbital changes to Saturn’s Phoebe ring and Kuiper belt dust take place. Given that these equations do not account for geometrical factors, these timescales should not be taken too precisely. These timescales are shown in Table 1. Note that they must be compared to the duration of the wind rather than the time since the event. For Saturn’s Phoebe ring, we assume a 10 μ𝜇\muitalic_μm ring particle with a density of 1.6 g/cm3 and semi-major axis 150 RSaturnsubscript𝑅SaturnR_{\rm Saturn}italic_R start_POSTSUBSCRIPT roman_Saturn end_POSTSUBSCRIPT (Verbiscer et al., 2009), whereas a Kuiper belt dust grain has a size of 1 μ𝜇\muitalic_μm and a density of 2 g/cm3 at 40 au. The SN density and velocity are calculated with a Sedov-Taylor profile in an ambient medium of 0.005 cm-3, taken from Miller & Fields (2022) and showing the onset of the blast. While the distance to the SN 3 Myr ago is uncertain, it is likely in the range of 20-140 pc based on Fe60superscriptFe60{}^{60}{\rm Fe}start_FLOATSUPERSCRIPT 60 end_FLOATSUPERSCRIPT roman_Fe measurements (Ertel & Fields, 2023; Fry et al., 2015). We include a 20 pc case here, representing an extreme event. In addition, since the incoming wind is completely governed by density and velocity, we include the scenario of passing through the LxCC, as proposed by Opher et al. (2024).

Table 1: Parameters for Ring Simulations Timescales for Orbit Perturbations
Case Wind density Wind velocity System a𝑎aitalic_a Timescale e𝑒eitalic_e Timescale i𝑖iitalic_i Timescale
n[atoms/cm3]𝑛delimited-[]atomssuperscriptcm3n\ [\rm atoms/cm^{3}]italic_n [ roman_atoms / roman_cm start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ] [km/s] τa[yr]subscript𝜏𝑎delimited-[]yr\tau_{a}\ [\rm yr]italic_τ start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT [ roman_yr ] τe[yr]subscript𝜏𝑒delimited-[]yr\tau_{e}\ [\rm yr]italic_τ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT [ roman_yr ] τi[yr]subscript𝜏𝑖delimited-[]yr\tau_{i}\ [\rm yr]italic_τ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT [ roman_yr ]
SN, 100 pc 0.02 273 Phoebe ring 2.8e7 4.3e5 8.7e5
SN, 100 pc 0.02 273 Kuiper Belt 1.8e6 6.1e4 1.2e5
SN, 50 pc 0.02 771 Phoebe ring 1.0e7 5.4e4 1.1e5
SN, 50 pc 0.02 771 Kuiper Belt 6.3e5 7.7e3 1.5e4
SN, 20 pc 0.02 3048 Phoebe ring 2.5e6 3.5e3 7.0e3
SN, 20 pc 0.02 3048 Kuiper Belt 1.6e5 490 980
LxCC 3000 14.1 Phoebe ring 3.7e3 1.1e3 2.2e3
LxCC 3000 14.1 Kuiper Belt 230 150 310

We find that the timescale for a SN to alter the orbit of particles in Saturn’s Phoebe ring is 104105superscript104superscript10510^{4}-10^{5}10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT years for the eccentricity and inclination, though the semi-major axis takes longer. A supernova blast must be closer than 50similar-toabsent50\sim 50∼ 50 pc to compress the heliosphere enough to reach Saturn in the first place (Miller & Fields, 2022; Tamayo et al., 2016). The exposure time would be in the range of 1-10 kyr, in surprising agreement with the eccentricity and inclination timescales calculated here. Therefore, we conclude that nearby supernovae may in fact affect the tilt and shape of the Phoebe ring. Passing through a dense interstellar cloud as in Opher et al. (2024) requires an exposure time of about 1 kyr. Given the uncertainties in the size of the LxCC, it is unclear whether or not this timescale could be achieved.

The results are not drastically different for Kuiper belt dust. But in this case, the SN influence occurs over 0.1-1 Myr, depending on the distance. When comparing with the exposure time from Miller & Fields (2022, their Fig. 11), we see that such timescales are typically around 50 kyr at a distance of 40 au, nearly independent of SN distance. A nearby SN may indeed alter the orbit of Kuiper belt dust grains. While the semi-major axis is not shifted greatly, the eccentricity and inclination have timescales two orders of magnitude less, and thus SNe will still modify their orbits in interesting ways. In their new orbits, the grains would experience additional effects like the gravitational tug of other planets.

The passage through dense interstellar clouds would certainly have a large effect of Kuiper belt dust. The orbit decay timescale τa200similar-tosubscript𝜏𝑎200\tau_{a}\sim 200italic_τ start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ∼ 200 yr for the LxCC is short, and so grains are significantly affected over even a single orbit. In the SN cases, τasubscript𝜏𝑎\tau_{a}italic_τ start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT was two orders of magnitude greater than τesubscript𝜏𝑒\tau_{e}italic_τ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT and τisubscript𝜏𝑖\tau_{i}italic_τ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. However, because the LxCC has such a lower velocity, eq. (11) implies that this difference in velocity is what causes the timescales to be similar here. Such an order of magnitude calculation suggests that this scenario deserves a closer calculation.

3.3 Numerical Integration

Given the surprisingly short timescales to modify the orbits of Kuiper belt dust, a numerical simulation is in order. The behavior of the particles is influenced both by the drag force created by the blast wave (eq. 2) and Newtonian gravity. A more complete model, such as those of e.g., Slavin et al. (2012) or Poppe (2016), would account for additional forces like magnetic fields or Poynting-Robertson drag. We leave such efforts for future work. The equation of motion for the particle of velocity u𝑢\vec{u}over→ start_ARG italic_u end_ARG is

mgrdudt=(1β)GMmgrr|r|3+Fdrag,subscript𝑚gr𝑑𝑢𝑑𝑡1𝛽𝐺subscript𝑀direct-productsubscript𝑚gr𝑟superscript𝑟3subscript𝐹dragm_{\mathrm{gr}}\frac{d\vec{u}}{dt}=-(1-\beta)GM_{\odot}m_{\mathrm{gr}}\frac{% \vec{r}}{|\vec{r}|^{3}}+F_{\rm drag},italic_m start_POSTSUBSCRIPT roman_gr end_POSTSUBSCRIPT divide start_ARG italic_d over→ start_ARG italic_u end_ARG end_ARG start_ARG italic_d italic_t end_ARG = - ( 1 - italic_β ) italic_G italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT italic_m start_POSTSUBSCRIPT roman_gr end_POSTSUBSCRIPT divide start_ARG over→ start_ARG italic_r end_ARG end_ARG start_ARG | over→ start_ARG italic_r end_ARG | start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG + italic_F start_POSTSUBSCRIPT roman_drag end_POSTSUBSCRIPT , (15)

which we integrate over several orbital periods. Here β𝛽\betaitalic_β is the ratio of radiation pressure force on the grain to gravity; β<1𝛽1\beta<1italic_β < 1 gives a net attractive force for grains that orbit the Sun in the absence of a wind. In our numerical results we will take β=0𝛽0\beta=0italic_β = 0, but for nonzero values the timescale would increase by a factor (1β)1/2superscript1𝛽12(1-\beta)^{-1/2}( 1 - italic_β ) start_POSTSUPERSCRIPT - 1 / 2 end_POSTSUPERSCRIPT.

As an example, Fig. 3 shows the orbit of a 1 μ𝜇\muitalic_μm dust grain subject to the high density of the LxCC (the “LxCC Kuiper Belt” case in Table 1). Initially, the grain has a semi-major axis of 40 au and an eccentricity of 0.1, and the wind blows in the [x^^𝑥\hat{x}over^ start_ARG italic_x end_ARG, y^^𝑦\hat{y}over^ start_ARG italic_y end_ARG, z^^𝑧\hat{z}over^ start_ARG italic_z end_ARG] direction, which was chosen to give an oblique angle to the ecliptic and to give bound trajectories. Over a duration of 500 years, the orbit changes drastically. The most apparent change is the semi-major axis, which has diminished to less than 1 au by the end. Also noteworthy is the inclination, which has been altered so greatly the particle orbits nearly perpendicular to its initial orbit. Significant changes occur on the order of 100similar-toabsent100\sim 100∼ 100 years, in good agreement with the secular approximation. In this instance, the dust grain remained bound in the solar system; however, by merely changing the direction of the wind, it is readily possible to eject the grain. For example, if the wind comes from the ecliptic pole, the grain is unbound from the system. For μ𝜇\muitalic_μm-sized dust, then, even the direction of the wind has a significant effect on the fate of Kuiper belt dust. But in all scenarios the grain no longer orbits in the Kuiper belt.

Refer to caption
Figure 3: One possible trajectory of a 1 μ𝜇\muitalic_μm dust grain in the Kuiper belt subject to a constant wind from a dense interstellar cloud, corresponding to the “LxCC” case in Table 1. (a) shows the full 3D path, and (b) and (c) show projections in the x𝑥xitalic_x-y𝑦yitalic_y and x𝑥xitalic_x-z𝑧zitalic_z planes, respectively. Note that (c) is not an equal aspect ratio.

This numerical exercise is repeated for two SN scenarios, a 20 pc and 50 pc case. They are given the same initial conditions as in the LxCC scenario, but are integrated over a longer time to be consistent with the duration of a SN passage (Miller & Fields, 2022). The evolution of the semi-major axes, eccentricities, and inclinations are shown in Fig. 4. As expected from the secular approximation, the eccentricity changes rapidly while the long-term evolution semi-major axis occurs over much longer timescales.

Refer to caption
Figure 4: Integrated time-evolution of a 1 μ𝜇\muitalic_μm Kuiper belt dust grain subject to a wind from a 20 pc SN (top), a 50 pc SN (middle), and the LxCC (bottom). Plotted quantities are the semi-major axis (blue solid), eccentricity (orange dashed), and inclination (green dotted). Note that in the bottom panel, the semi-major axis and time are on different scales than the top and middle.

These numerical simulations, in addition to validating the timescale estimates of the secular approximation, show exactly how dust grain orbits react to external winds. Further study is needed to apply this process to the observed distribution of interplanetary dust, which will show how the entire dust distribution changes.

4 Discussion & Conclusions

We investigated how a nearby SN blast can dynamically alter the orbits of bodies in our solar system. In the first case, we employed an impulse approximation to model the effects on Oort Cloud bodies. By allowing for a wide range of radii, we found that a SN 50 pc distant will leave most 10greater-than-or-equivalent-toabsent10\gtrsim 10≳ 10 cm bodies unaltered in their orbit, but ejects all dust grains smaller than 1 mm from the solar system. From the geological Fe60superscriptFe60{}^{60}{\rm Fe}start_FLOATSUPERSCRIPT 60 end_FLOATSUPERSCRIPT roman_Fe-based rate of 2similar-toabsent2\sim 2∼ 2 SNe per 10 Myr, these events would sweep clear the Oort Cloud of small debris many times over the age of the solar system. Shortly after the event, debris will be replaced as collisions occur and larger bodies erode. Stern & Shull (1988) has also shown that another effect of nearby SNe is to flash-heat icy bodies, which could potentially alter their surface properties.

In addition to our own solar system, a roughly spherical SN blast will also collide with all nearby stellar systems. With a similar stellar density to the solar neighborhood, the nearest of these stars will be similar-to\sim1 pc distant. If that star hosts an exo-Oort Cloud, then the supernova blast could eject objects of up to 100similar-toabsent100\sim 100∼ 100 m into interstellar space, seeding the Milky Way with interstellar comets and asteroids. Thus, SN ejection represents a possible mechanism to create objects like ’Oumuamua (Meech et al., 2017).

A secular approximation is applied to the orbits of Saturn’s Phoebe ring and the Kuiper belt. The timescale for a SN to significantly alter Saturn’s A and B rings is extremely large, but that of the Phoebe ring is surprisingly within reach for a SN within 50 pc. For a dense interstellar cloud, this timescale is on the order of similar-to\simkyr. If our solar system crossed the LxCC, it would have had a minor effect on the Phoebe ring. However, crossings of GMCs typically take 0.1-1 Myr (Talbot & Newman, 1977), and so would certainly devastate the Phoebe ring. While crossing GMCs with a density high enough to expose Earth statistically occur every similar-to\simGyr, the density required to expose Saturn (assuming ram pressure balance) is (aS/a)2=90superscriptsubscript𝑎Ssubscript𝑎direct-sum290(a_{\rm S}/a_{\oplus})^{2}=90( italic_a start_POSTSUBSCRIPT roman_S end_POSTSUBSCRIPT / italic_a start_POSTSUBSCRIPT ⊕ end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 90 times less, and should occur much more frequently.

Based on the same secular approximation, the distribution of Kuiper belt dust exposed to a SN blast may be significantly altered. If the ejection fraction were roughly similar to that of the Oort Cloud, a supernova 50 pc distant (such as the one 3 Myr ago) would have ejected all particles less than similar-to\sim1 mm from the Kuiper belt. This result would indicate that the dust observed by New Horizons (Poppe et al., 2019) has been generated in the past 3 Myr. Indeed, such dust may not have reached an equilibrium state, as the Poynting-Robertson drag time is similar-to\sim5 Myr (Moro-Martín & Malhotra, 2003; Wyatt & Whipple, 1950). In addition, the observed Kuiper belt dust production rate of 107superscript10710^{7}10 start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT g s-1 (Poppe, 2016) and a total dust mass of 3.5×10183.5superscript10183.5\times 10^{18}3.5 × 10 start_POSTSUPERSCRIPT 18 end_POSTSUPERSCRIPT kg (Poppe et al., 2019) lead to a dust “replenishment” timescale of 11 Myr. Both these Poynting-Robertson and replenishment timescales are longer than the age of the most recent SN blast, so it is possible that echoing effects still exist in dust grain trajectories today. Furthermore, recent reports from the New Horizons spacecraft have shown an unexpected increase in dust flux (Doner et al., 2024). Whether this increase in dust flux is in some way related to recent interstellar factors is left for future study. These efforts may determine whether extant effects in Kuiper belt dust may bear the scars of ancient SNe.

We gladly acknowledge many enlightening discussions with Adrienne Ertel, Erica Albrigo, Merav Opher, and Alex Doner. We are particularly indebted to Scott Tremaine for discussions which inspired this work. We also thank the Undergraduate Research Apprenticeship Program (URAP) at UIUC. JAM acknowledges support from the NASA/DRIVE program entitled “Our Heliospheric Shield”, 80NSSC22M0164, https://shielddrivecenter.com. The work of BDF was supported by the NSF under grant number AST-210858.

References

  • Athanassiadou & Fields (2011) Athanassiadou, T., & Fields, B. D. 2011, New Astronomy, 16, 229
  • Doner et al. (2024) Doner, A., Horányi, M., Bagenal, F., et al. 2024, ApJL, 961, L38
  • Dones et al. (2015) Dones, L., Brasser, R., Kaib, N., & Rickman, H. 2015, Space Sci Rev, 197, 191
  • Ellis et al. (1996) Ellis, J., Fields, B. D., & Schramm, D. N. 1996, The Astrophysical Journal, 470, 1227
  • Ertel & Fields (2023) Ertel, A. F., & Fields, B. D. 2023, Distances to Recent Near-Earth Supernovae From Geological and Lunar 60Fe, arXiv. https://arxiv.org/abs/2309.11604
  • Ertel et al. (2023) Ertel, A. F., Fry, B. J., Fields, B. D., & Ellis, J. 2023, ApJ, 947, 58
  • Fitoussi et al. (2008) Fitoussi, C., Raisbeck, G. M., Knie, K., et al. 2008, Phys. Rev. Lett., 101, 121101
  • Fry et al. (2015) Fry, B. J., Fields, B. D., & Ellis, J. R. 2015, ApJ, 800, 71
  • Fry et al. (2020) —. 2020, ApJ, 894, 109
  • Gehrels et al. (2003) Gehrels, N., Laird, C. M., Jackman, C. H., et al. 2003, ApJ, 585, 1169
  • Harris et al. (2020) Harris, C. R., Millman, K. J., van der Walt, S. J., et al. 2020, Nature, 585, 357
  • Heisler & Tremaine (1986) Heisler, J., & Tremaine, S. 1986, Icarus, 65, 13
  • Hunter (2007) Hunter, J. D. 2007, Computing in Science Engineering, 9, 90
  • Hut & Tremaine (1985) Hut, P., & Tremaine, S. 1985, The Astronomical Journal, 90, 1548
  • Jakubík & Neslušan (2008) Jakubík, M., & Neslušan, L. 2008, Contributions of the Astronomical Observatory Skalnate Pleso, 38, 33
  • Knie et al. (2004) Knie, K., Korschinek, G., Faestermann, T., et al. 2004, Physical Review Letters, 93
  • Knie et al. (1999) —. 1999, Physical Review Letters, 83, 18
  • Ludwig et al. (2016) Ludwig, P., Bishop, S., Egli, R., et al. 2016, Proceedings of the National Academy of Sciences, 113, 9232
  • Meech et al. (2017) Meech, K. J., Weryk, R., Micheli, M., et al. 2017, Nature, 552, 378
  • Miller & Fields (2022) Miller, J. A., & Fields, B. D. 2022, The Astrophysical Journal, 934, 32
  • Moro-Martín & Malhotra (2003) Moro-Martín, A., & Malhotra, R. 2003, AJ, 125, 2255
  • Oort (1950) Oort, J. H. 1950, Bulletin of the Astronomical Institutes of the Netherlands, 11, 91
  • Opher et al. (2024) Opher, M., Loeb, A., & Peek, J. E. G. 2024, Nat Astron, 1
  • Pástor et al. (2011) Pástor, P., Klačka, J., & Kómar, L. 2011, Monthly Notices of the Royal Astronomical Society, 415, 2637
  • Poppe (2016) Poppe, A. R. 2016, Icarus, 264, 369
  • Poppe et al. (2019) Poppe, A. R., Lisse, C. M., Piquette, M., et al. 2019, ApJL, 881, L12
  • Price-Whelan et al. (2018) Price-Whelan, A. M., Sipőcz, B. M., Günther, H. M., et al. 2018, AJ, 156, 123
  • Rickman et al. (2008) Rickman, H., Fouchard, M., Froeschlé, C., & Valsecchi, G. B. 2008, Celestial Mechanics and Dynamical Astronomy, 102, 111
  • Robitaille et al. (2013) Robitaille, T. P., Tollerud, E. J., Greenfield, P., et al. 2013, A&A, 558, A33
  • Slavin et al. (2012) Slavin, J. D., Frisch, P. C., Müller, H.-R., et al. 2012, ApJ, 760, 46
  • Stern (1990) Stern, S. A. 1990, Icarus, 84, 447
  • Stern & Shull (1988) Stern, S. A., & Shull, J. M. 1988, Nature, 332, 407
  • Talbot & Newman (1977) Talbot, Jr., R. J., & Newman, M. J. 1977, The Astrophysical Journal Supplement Series, 34, 295
  • Tamayo et al. (2016) Tamayo, D., Markham, S. R., Hedman, M. M., Burns, J. A., & Hamilton, D. P. 2016, Icarus, 275, 117
  • Thomas et al. (2016) Thomas, B. C., Engler, E. E., Kachelrieß, M., et al. 2016, The Astrophysical Journal, 826, L3
  • Verbiscer et al. (2009) Verbiscer, A. J., Skrutskie, M. F., & Hamilton, D. P. 2009, Nature, 461, 1098
  • Virtanen et al. (2020) Virtanen, P., Gommers, R., Oliphant, T. E., et al. 2020, Nat Methods, 17, 261
  • Wallner et al. (2016) Wallner, A., Feige, J., Kinoshita, N., et al. 2016, Nature, 532, 69
  • Wallner et al. (2021) Wallner, A., Froehlich, M. B., Hotchkis, M. a. C., et al. 2021, Science, 372, 742
  • Wyatt & Whipple (1950) Wyatt, S. P., & Whipple, F. L. 1950, The Astrophysical Journal, 111, 134