www.fgks.org   »   [go: up one dir, main page]

SlideShare a Scribd company logo
The Limited Role of the Streaming Instability during Moon and Exomoon Formation
Miki Nakajima1,2
, Jeremy Atkins2
, Jacob B. Simon3
, and Alice C. Quillen2
1
Department of Earth and Environmental Sciences, University of Rochester, P.O. Box 270221, Rochester, NY 14627, USA
2
Department of Physics and Astronomy, University of Rochester, P.O. Box 270171, Rochester, NY 14627, USA
3
Department of Physics and Astronomy, Iowa State University, Ames, IA 50010, USA
Received 2023 September 20; revised 2024 March 27; accepted 2024 April 17; published 2024 June 17
Abstract
It is generally accepted that the Moon accreted from the disk formed by an impact between the proto-Earth and
impactor, but its details are highly debated. Some models suggest that a Mars-sized impactor formed a silicate
melt-rich (vapor-poor) disk around Earth, whereas other models suggest that a highly energetic impact produced a
silicate vapor-rich disk. Such a vapor-rich disk, however, may not be suitable for the Moon formation, because
moonlets, building blocks of the Moon, of 100 m–100 km in radius may experience strong gas drag and fall onto
Earth on a short timescale, failing to grow further. This problem may be avoided if large moonlets (?100 km)
form very quickly by streaming instability, which is a process to concentrate particles enough to cause gravitational
collapse and rapid formation of planetesimals or moonlets. Here, we investigate the effect of the streaming
instability in the Moon-forming disk for the first time and find that this instability can quickly form ∼100 km-sized
moonlets. However, these moonlets are not large enough to avoid strong drag, and they still fall onto Earth quickly.
This suggests that the vapor-rich disks may not form the large Moon, and therefore the models that produce vapor-
poor disks are supported. This result is applicable to general impact-induced moon-forming disks, supporting the
previous suggestion that small planets (<1.6 R⊕) are good candidates to host large moons because their impact-
induced disks would likely be vapor-poor. We find a limited role of streaming instability in satellite formation in an
impact-induced disk, whereas it plays a key role during planet formation.
Unified Astronomy Thesaurus concepts: Earth-moon system (436)
Supporting material: tar.gz file
1. Introduction
1.1. Review of the Moon-formation Hypotheses
The origin of the Earth’s Moon is a long-standing open
problem in planetary science. While it is accepted that the
Moon formed from a partially vaporized disk generated by a
collision between the proto-Earth and an impactor approxi-
mately 4.5 billion yr ago (the “giant impact hypothesis”;
Hartmann & Davis 1975; Cameron & Ward 1976), the details,
such as the impactor radius and velocity, are actively debated.
According to the canonical model, the proto-Earth was hit by a
Mars-sized impactor (Canup & Asphaug 2001). This model can
successfully explain several observations, such as the mass of
the Moon, the angular momentum of the Earth–Moon system,
and the small lunar core. Moreover, this model could
potentially explain the observation that the Moon is depleted
in volatiles, if they escaped during the impact or subsequent
processes (see reviews by Halliday 2004; Canup et al. 2023).
However, this model fails to explain the Earth and Moon’s
nearly identical isotopic ratios (e.g., Si, Ti, W, and O;
Armytage et al. 2012; Zhang & Dauphas 2012; Touboul
et al. 2015; Thiemens et al. 2019, 2021; Kruijer et al. 2021).
This is because the disk generated by an oblique Mars-sized
impactor is primarily made of the impactor materials
(Canup 2004; Kegerreis et al. 2022; Hull et al. 2024), which
likely had different isotopic ratios from Earth, unless the inner
solar system was well mixed and homogenized (Dauphas 2017)
or equilibrated in the disk phase (Pahlevan & Stevenson 2007;
Lock et al. 2018). In contrast, more energetic impact models
may solve this problem by mixing the proto-Earth and
impactor. These energetic models include an impact between
two half-Earth-sized objects (Canup 2012), an impact between
a rapidly rotating proto-Earth and a small impactor (Cúk &
Stewart 2012), and general impact events that involve high
energy and high angular momentum, which create a doughnut-
shaped vapor disk connected to the Earth (a so-called synestia;
Lock et al. 2018). In these scenarios, the disk and the proto-
Earth compositions could have been mixed well, potentially
solving the isotopic problem. Other proposed scenarios include
(a) the Moon formation by multiple small impacts (Rufu et al.
2017), (b) hit-and-run collisions that mix the proto-Earth and
the Moon more than the canonical model does (Asphaug et al.
2021), and (c) an impact between the proto-Earth covered by a
surface magma ocean and an impactor, which leads to more
contribution from Earth to the disk than the canonical scenario
(Hosono et al. 2019).
All of these models have potential explanations for the
isotopic similarity, but each model faces challenges to explain
other constraints. For example, the half-Earth and synestia
models predict a much higher angular momentum of the Earth–
Moon system than that of today (Canup 2012; Cúk &
Stewart 2012). The excess of the angular momentum may or
may not be removed easily (Cúk & Stewart 2012; Cúk et al.
2016, 2021; Rufu et al. 2017; Rufu & Canup 2020; Ward et al.
2020). Another constraint may come from the Earth’s interior;
the deep portion of the Earth still retains solar neon (Yokochi &
Marty 2004; Williams & Mukhopadhyay 2019) and helium
isotopic ratios (Williams et al. 2019), which are distinct from
the rest of the mantle. These noble gases may have been
The Planetary Science Journal, 5:145 (12pp), 2024 June https://doi.org/10.3847/PSJ/ad4863
© 2024. The Author(s). Published by the American Astronomical Society.
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
delivered to Earth’s mantle during the planet formation phase.
This suggests that Earth’s mantle developed chemical hetero-
geneity during Earth’s accretion, when the protoplanetary disk
was still present, and that the Earth’s mantle has never been
completely mixed, even by the Moon-forming impact. Previous
work suggests that the preservation of the mantle heterogeneity
can be explained by the canonical model but not by the
energetic models, because these energetic impacts tend to mix
the mantle (Nakajima & Stevenson 2015). It should be noted,
however, the possibility that the neon and helium were
delivered to the Earth's core and have been slowly leaking
into the mantle over time has been discussed (Bouhifd et al.
2020), which would eliminate the need to preserve the mantle
heterogeneity. However, it is not clear if delivery of these noble
gases to the core was efficient.
Moreover, due to recent analytical capability, small isotopic
differences between Earth and the Moon have been observed
(e.g., K, O, W, V, Cr, and others; Wiechert et al. 2001; Kruijer
et al. 2015; Touboul et al. 2015; Wang & Jacobsen 2016;
Young et al. 2016; Thiemens et al. 2019; Nielsen et al. 2021;
Sossi et al. 2018). Some isotopic difference, such as the
Moon’s enrichment in heavy K isotopes compared to those of
Earth, can be explained by isotopic fractionation due to liquid–
vapor phase separation during the Moon accretion process
(Wang & Jacobsen 2016; Nie & Dauphas 2019; Charnoz et al.
2021). Observed small oxygen isotopic differences between
Earth and the Moon may suggest that the Moon still records the
impactor component (Cano et al. 2020). This suggestion may
not be compatible with the energetic impact models because
these impacts are so energetic that Earth and the Moon would
be efficiently mixed. Whether the proposed models for the
lunar origin can explain the observation that the Moon is
depleted in volatiles is actively debated (Canup et al. 2015,
2023; Dauphas et al. 2015, 2022; Nakajima & Stevenson 2018;
Sossi et al. 2018; Nie & Dauphas 2019; Charnoz et al. 2021;
Halliday & Canup 2022; Canup et al. 2023).
1.2. Gas Drag Problem in a Vapor-rich Disk
Another key constraint that has not been discussed until
recently is the vapor mass fraction (VMF) of the disk, which
sensitively depends on the impact model. Less energetic
impacts, such as the canonical model and the multiple impact
model, produce relatively small VMFs (VMF ∼ 0.2 for the
canonical model, Nakajima & Stevenson 2014, and ∼0.1–0.5
for the multiple impact model, Rufu et al. 2017), while more
energetic models, such as the half-Earth model and synestia
model, produce nearly pure vapor disks (VMF ∼0.8–1,
Nakajima & Stevenson 2014). The VMF of the Moon-forming
disk can significantly impact the Moon accretion process; if the
initial Moon-forming disk is vapor-poor (liquid-rich), moonlets
can quickly form by gravitational instability (GI) from the
liquid layer in the disk midplane outside the Roche radius
(Thompson & Stevenson 1988; Salmon & Canup 2012). These
moonlets eventually accrete to the Moon in tens to hundreds of
years (Thompson & Stevenson 1988; Salmon & Canup 2012;
Lock et al. 2018). In contrast, an initially vapor-rich disk needs
to cool until liquid droplets emerge before moonlet accretion
begins. Growing moonlets in the vapor-rich disk experience
strong gas drag from the vapor (Nakajima et al. 2022). The gas
drag effect is strongest when the gas and moonlet are coupled
to an adequate degree and is sensitive to the moonlet radius. It
is strongest with a moonlet radius Rp of a few kilometers
(Nakajima et al. 2022). In contrast, much smaller moonlets are
completely coupled with the gas, and much larger moonlets are
completely decoupled from the gas, experiencing a weaker gas
drag effect. As a result, ∼kilometer-sized moonlets lose their
angular momentum and inspiral onto the Earth within 1 day
(Nakajima et al. 2022), a much shorter timescale than that
required for lunar formation.
This same problem was a major challenge to planet
formation in the protoplanetary disk (Adachi et al. 1976;
Weidenschilling 1977). Here, we quickly review the gas drag
problem in the protoplanetary disk. The radial velocity of a
particle in a disk is (see Equation (A6) for the derivation; see
also Armitage 2010; Takeuchi & Lin 2002)
( )
v
v v
2
, 1
r
r
f
1
,g K
f f
1
t h
t t
=
-
+
-
-
where τf = Ωtf is the dimensionless stopping time (Armitage 2010),
Ω is the Keplerian angular velocity, tf is the friction time
(see further descriptions below), vr,g is the gas velocity, and η is the
pressure gradient parameter, described as
⎜ ⎟
⎛
⎝
⎞
⎠
( )
c
v
p
r
1
2
ln
ln
, 2
s
K
2
g
h = -
¶
¶
where cs is the sound speed, vK is the Keplerian velocity, pg is
the gas pressure, and r is the radial distance. vr is largest when
τf = 1. The friction time is the time until the particle and gas
reach the same velocity and is defined as tf = mvrel/FD, where
m is the particle mass, vrel is the relative azimuthal velocity
between the gas and particle, and FD is the drag force. In the
protoplanetary disk, the gas drag for a particle is generally
written as F R v
C
D 2 g,0 p
2
rel
2
D
pr
= - , where CD is the gas drag
coefficient, ρg,0 is the initial gas density at the midplane, and Rp
is the particle radius. The gas drag coefficients are roughly
(Armitage 2010)
⎧
⎨
⎩
( )
( )
( )
( )
C
24 Re , Re 1 Stokesregime
24 Re , 1 Re 800 transitionregime
0.44, Re 800 Newtonregime
, 3
D
1
0.6
=
<
< <
>
-
-
where Re is the Reynolds number. Assuming that vrel ∼ ηvK
(see Equation (A2)), ν ∼ csλ, where ν is the kinematic
viscosity, λ is the mean free path (
k T
d p
2
B
2
g,0
l =
p
, where kB is
the Boltzmann constant, T is the temperature, d is the molecular
diameter, and pg,0 is the gas pressure at the midplane), Re =
4.26
v Rp
rel
~
n
, which is in the transition regime, in the
protoplanetary disk at 1 au, assuming T = 280 K, η = 0.002,
d = 289 pm, Rp = 0.52 m (see discussion below), p
RT
m
g,0
g,0
m
=
r
,
and mm = 0.002 kg mol–1
, where mm is the mean molecular
weight and R is the gas constant. Here we also assume the
following vertical distribution of the gas density ρg,
⎜ ⎟
⎛
⎝
⎞
⎠
( )
z
H
exp
2
, 4
g g,0
2
2
r r
= -
where z is the vertical coordinate and H is the gas scale height.
This leads to
H
g,0 2
g
r =
p
S
. We assume Σg = 17,000 kg m−2
and
use the relationships of c RT m
s m
= , H = cs/Ω,
p
r
ln
ln
g
=
¶
¶
2
The Planetary Science Journal, 5:145 (12pp), 2024 June Nakajima et al.
( )
3 2
b
- + , and
1
2
b = . The orbital angular frequency is
GM r3
W = , where G is the gravitational constant, M is the
stellar mass and r is the distance from the star (for the Moon-
forming disk, M is the Earth mass and r is the distance from
Earth). These parameters are taken from previous work
(Carrera et al. 2015; other values of β have been discussed,
such as
3
7
b = ; Chiang & Youdin 2010).
The dimensionless stopping time becomes
( )
t
C
R
v C
R
r
8
3
8
3
. 5
f f
D
p p
g,0 rel D
p p
g,0
t
r
r h
r
r
= W = W =
τf = 1 when Rp = 0.52 m. The particle density ρp = 3000 kg
m−3
is assumed. The residence time of the particle at 1 au is
1 au/vr = 75 yr, where vr is the radial fall velocity of the
particle (see Equation (A7); for simplicity, the radial gas
velocity vr,g = 0 is assumed). Thus, an approximately meter-
sized particle falls toward the Sun at 1 au within 80 yr, a
timescale much shorter than the planet formation timescale
(several tens of Myr). This is the so-called “meter barrier”
problem, which was a major issue to explain planetary growth
in a classical planet formation model (Adachi et al. 1976;
Weidenschilling 1977).
In contrast, for the vapor-rich Moon-forming disk, τf = 1
(Re 1010
> ) is achieved when Rp = 1.8 km, assuming
ρg,0 = 40 kg m–3
, ρp = 3000 kg m–3
, r = 3 R⊕, η = 0.04,
T = 5200 K (see Section 2.3 for justifications for these parameters),
and d = 300 pm. The residence time, r/vr, is approximately 1 day
(1.21 days), which is much shorter than the Moon-formation
timescale of several tens of yr to 100 yr. This formation timescale
is ultimately determined by the radiative cooling timescale, but it is
model-dependent. Here we provide a very simple estimate; the
timescale for radiative cooling is
( )
10
M L C T
r T
4
p
disk
2
ph
4 ~
p s
+ D
yr (this is
also consistent with numerical work; Lock et al. 2018), where
Mdisk is the disk mass, L is the latent heat, Cp is the specific heat,
ΔT is the temperature change over time, σ is the Stefan–
Boltzmann constant, and Tph is the photosphere temperature. Here,
Mdisk ∼ 0.015 M⊕, L = 1.2 × 107
J kg–1
(Melosh 2007), Cp =
103
J K–1
kg–1
, ΔT = 2000 K, and Tph = 2000 K (Thompson &
Stevenson 1988) are assumed. However, the actual Moon-
formation timescale can be longer than this for several reasons,
including additional heating due to viscous spreading (Thompson
& Stevenson 1988; Charnoz & Michaut 2015) and radial material
transport efficiency (e.g., Salmon & Canup 2012). A short Moon-
formation timescale has been proposed (e.g., Mullen & Gam-
mie 2020; Kegerreis et al. 2022), but the typical Moon-formation
timescale has been estimated to range in 10–102
yr.
The vertical settling velocity of condensing particles is
v C
R z
settle
8
3 D
p p
2
g
=
r
r
W
(Armitage 2010). Assuming z ∼ H, the
settling time for a particle with a radius off 2 is 0.22 days.
Determining the collision history of moonlets requires
conducting orbital dynamics simulations, but here we provide
a rough estimate. A rough estimate of the mass-doubling time
of the largest moonlet in the Moon-forming disk is typically
∼1 day (Salmon & Canup 2012). If a 2 km-sized moonlet mass
doubles in a day, the radius change is very small (2 km ×
21/3
= 2.5 km), and the gas drag effect remains strong on such
a moonlet; this change does not prevent the moonlet from
falling into Earth. However, the actual collision time can vary
depending on the local concentration of particles, which needs
a detailed future study.
This gas drag effect is a problem with forming the Moon
from an initially vapor-rich disk. The Moon can still form after
most of the vapor condenses, but by that time a significant
portion of the disk mass could be lost (Ida et al. 2020;
Nakajima et al. 2022), which would fail to form a large moon
from an initially vapor-rich disk (here we use a ”large” moon
when its mass is ∼1 wt% or larger of the host planet). If this is
the case, an initially vapor-rich disk may not be capable of
forming a large Moon (Nakajima et al. 2022). In contrast, the
gas drag effect is weak for vapor-poor disks, which are
generated by less energetic models, such as the canonical and
multiple impact models.
1.3. Streaming Instability in the General Impact-induced
Moon-forming Disk
A potential solution to this vapor drag problem is forming a
large moonlet very quickly (much larger than 2 km), so that the
moonlet would not experience strong gas drag. This is the
accepted solution for the gas drag problem in the protoplane-
tary disk. The proposed mechanism is the streaming instability
(Youdin & Goodman 2005; Johansen et al. 2007), where
particles spontaneously concentrate in the disk, gravitationally
collapsing and forming a large clump (∼100 km in size;
Johansen et al. 2015). If this mechanism works for the Moon-
forming disk, an initially vapor-rich disk may be able to form
the Moon despite the gas drag issue. If this mechanism turns
out not to work for the Moon-forming disk, it is an interesting
finding as well, given that a Moon-forming disk is often treated
as a miniature analog of the protoplanetary disk. Understanding
what makes these two disks differ would deepen our under-
standing of planet and satellite formation processes.
Moreover, knowing whether the streaming instability can affect
moon formation processes informs our understanding of moon
formation in the solar and extrasolar systems. Moon formation in
an impact-induced disk is common in the solar system (e.g., the
moons of Mars (Craddock 2011), the moons of Uranus (Slattery
et al. 1992), and Pluto-Charon (Canup 2005)). While there are no
confirmed exomoons (moons around exoplanets) to date (e.g.,
Cassese & Kipping 2022), impact-induced exomoons should be
common because impacts are a common process during planet
formation (Nakajima et al. 2022) and because impacts in
extrasolar systems may have already been observed (e.g., Meng
et al. 2014; Bonomo et al. 2019; Thompson et al. 2019;
Kenworthy et al. 2023). If streaming instability operates in these
disks, it can affect what types of planets can host exomoons,
which can be compared with future exomoon observations.
It should also be noted that other instability, such as secular
GI (e.g., Youdin 2011; Takahashi & Inutsuka 2014; Tominaga
et al. 2018) and two-component viscous GI (TVGI; Tominaga
et al. 2019), have been discussed as a mechanism for
planetesimal and dust ring formation. The secular GI occurs
when dust–gas interaction reduces the rotational support of the
rotating disk, which leads to dust concentration. The TVGI is
an instability caused by dust–gas friction and turbulent gas
viscosity. These instabilities can lead to clump formation even
if the disk is self-gravitationally stable. The implications of
these instabilities are beyond the scope of this paper.
3
The Planetary Science Journal, 5:145 (12pp), 2024 June Nakajima et al.
1.4. Motivation of This Work
The goal of this work is to investigate whether the streaming
instability can form moonlets that are large enough to avoid
strong gas drag from the vapor-rich disk. The result could
constrain the Moon-formation model as well as general impact-
induced models in the solar and extrasolar systems. We
conduct hydrodynamic simulations using the code Athena
(Stone et al. 2008; Bai & Stone 2010a). We first conduct 2D
simulations to identify the section of parameter space that leads
to streaming instability. Subsequently, we conduct 3D simula-
tions with self-gravity to identify the size distribution of the
moonlets. Finally, we identify the lifetime of the moonlets
formed by streaming instability to investigate whether it is
possible to form a large moon from an initially vapor-rich disk.
For the general impact-induced disks, we consider ”rocky” and
”icy” disks, where these disks form by collisions between
rocky planets (with silicate mantles and iron cores) and
between icy planets (with water-ice mantles and silicate cores),
respectively.
2. Method
2.1. Athena
We use the Athena hydrodynamics code, which solves the
equations of hydrodynamics using a second-order accurate
Godunov flux-conservative approach (Stone et al. 2008). We
use the configuration of Athena that couples the dimensionally
unsplit corner transport upwind method (Colella 1990) to the
third-order in-space piecewise parabolic method by Colella &
Woodward (1984) and calculates the numerical fluxes using the
HLLC Riemann solver (Toro 1999). We also integrate the
equations of motion for particles following Bai & Stone
(2010a) and include particle self-gravity for 3D simulations
following the particle-mesh approach described in previous
work (Simon et al. 2016). Orbital advection is taken into
account following previous work (Bai & Stone 2010a, 2010b).
Our setup is the local shearing box approximation in which a
small patch of the disk is corotating with the disk at the
Keplerian velocity (Stone & Gardiner 2010). The local
Cartesian frame is defined as (x, z) for 2D and (x, y, z) for
3D simulations, with x as the radial coordinate from the planet,
z parallel to the planetary spin axis, and y in the direction of the
orbital rotation.
In the Athena code, we solve the following equations for our
simulations (Bai & Stone 2010a; Simon et al. 2016; Li et al.
2019):
· ( ) ( )
t
u 0, 6
g
g
r
r
¶
¶
+  =
· ( )
( )
t
p
t
u
uu I
x z u
v u
3 2 , 7
g
g
g
g
2
g
2
g p
f
r
r
r r r r
W
¶
¶
+  +
= W - W - ´ +
-
ˆ ˆ
ˆ ( )
d
dt
v x
z
t
v
x x
z v
v u
a
2 3
2 . 8
i
i
i i
i
K
2
2
f
g
h
W
= - W + W
-W - ´ -
-
+
The first two equations specify mass and momentum
conservation for the gas, respectively, while the third equation
represents the motion of a particle i coupled with the gas. Here,
u is the velocity of the gas and I is the identity matrix. v is the
mass-weighted averaged particle velocity in the fluid element,
assuming that particles can be treated as fluid (Bai &
Stone 2010a). The terms of the right-hand side of
Equation (7) are radial tidal forces (gravity and centrifugal
force), vertical gravity, and the Coriolis force and the feedback
from the particle to the gas. In Equation (8), the first term on
the right-hand side describes a constant radial force due to gas
drag. vi is the particle velocity, x̂ and ẑ represent the unit
vectors in the x- and z-directions, and xi and zi are the values of
x and z for the particle i. u is the gas velocity interpolated from
the grid cell centers to the location of the particle. The second,
third, and fourth terms are radial and vertical tidal forces and
the Coriolis force. ag is the acceleration due to self-gravity,
which is considered only in 3D simulations. ag = −∇Φp, where
Φp is the gravitational potential and satisfies the Poisson’s
equation, ∇2
Φp = 4πGρp. To reduce computational time, the
particles are organized into “superparticles,” each representing
a cluster of individual particles of the same size. In the code
units, we normalize Ω = cs = H = ρg,0 = 1. The gas and
particle initial distributions are described in Equation (4) and as
⎛
⎝
⎜
⎞
⎠
⎟ ( )
H
z
H
2
exp
2
, 9
p
p
p
2
p
2
r
p
=
S
-
respectively, where Hp is the scale height of the particles and is
set to 0.02H (Bai & Stone 2010a) and Σp is the particle surface
density. The system uses an isothermal equation of state
P c
g s
2
r
= , and the particles are distributed uniformly in the x-
and y-directions and normally in the z-direction (Equation (9)).
The computational domains are 0.2H × 0.2H in 2D simulations
and 0.2H × 0.2H × 0.2H in 3D simulations, with all-periodic
boundary conditions. The resolution for 2D is 512 × 512, and
each grid cell has one particle (512 × 512 = 262,144 particles).
The resolution for 3D is 10 × 1283
(= 21 million particles in
total). Previous work shows that the resolution of 1283
can
produce the large clump size distribution well compared to
those with 2563
and 5123
(Simon et al. 2016); therefore, this
resolution of our 3D simulations is sufficient to resolve large
clumps, which are the focus of this work. The initial particle
size is assumed to be constant. Variable initial particle sizes
could affect the growth speed (Krapp et al. 2019) and the
concentrations of particles depending on their sizes (Yang &
Zhu 2021), but it is not known to affect the largest clump size
formed by streaming instability.
2.2. Clump Detection in 2D and 3D
After we conduct 2D simulations, we identify filaments
forming in the disk in order to constrain the parameter space
favorable for the streaming instability. We use the Kolmo-
gorov–Smirnov (KS) method, which is based on previous work
(Carrera et al. 2015). For each time step in a given time
window (25/Ω in our case), we compute the particle surface
density ( ) ( )
x x z dz
,
p p
0.1
0.1
ò r
S =
-
and average it over the time
window to give p t
áS ñ . We then sort the values in p t
áS ñ from
highest to lowest and compute the cumulative distribution
of this sorted data set. Finally, we use the KS test to output a
4
The Planetary Science Journal, 5:145 (12pp), 2024 June Nakajima et al.

Recommended for you

Formation of low mass protostars and their circumstellar disks
Formation of low mass protostars and their circumstellar disksFormation of low mass protostars and their circumstellar disks
Formation of low mass protostars and their circumstellar disks

Understanding circumstellar disks is of prime importance in astrophysics, however, their birth process remains poorly constrained due to observational and numerical challenges. Recent numerical works have shown that the small-scale physics, often wrapped into a sub-grid model, play a crucial role in disk formation and evolution. This calls for a combined approach in which both the protostar and circumstellar disk are studied in concert. Aims. We aim to elucidate the small scale physics and constrain sub-grid parameters commonly chosen in the literature by resolving the star-disk interaction. Methods. We carry out a set of very high resolution 3D radiative-hydrodynamics simulations that self-consistently describe the collapse of a turbulent dense molecular cloud core to stellar densities. We study the birth of the protostar, the circumstellar disk, and its early evolution (< 6 yr after protostellar formation). Results. Following the second gravitational collapse, the nascent protostar quickly reaches breakup velocity and sheds its surface material, thus forming a hot (∼ 103 K), dense, and highly flared circumstellar disk. The protostar is embedded within the disk, such that material can flow without crossing any shock fronts. The circumstellar disk mass quickly exceeds that of the protostar, and its kinematics are dominated by self-gravity. Accretion onto the disk is highly anisotropic, and accretion onto the protostar mainly occurs through material that slides on the disk surface. The polar mass flux is negligible in comparison. The radiative behavior also displays a strong anisotropy, as the polar accretion shock is shown to be supercritical whereas its equatorial counterpart is subcritical. We also f ind a remarkable convergence of our results with respect to initial conditions. Conclusions. These results reveal the structure and kinematics in the smallest spatial scales relevant to protostellar and circumstellar disk evolution. They can be used to describe accretion onto regions commonly described by sub-grid models in simulations studying larger scale physics.

protostarscircunstellar disksstar formation
The debris of the ‘last major merger’ is dynamically young
The debris of the ‘last major merger’ is dynamically youngThe debris of the ‘last major merger’ is dynamically young
The debris of the ‘last major merger’ is dynamically young

The Milky Way’s (MW) inner stellar halo contains an [Fe/H]-rich component with highly eccentric orbits, often referred to as the ‘last major merger.’ Hypotheses for the origin of this component include Gaia-Sausage/Enceladus (GSE), where the progenitor collided with the MW proto-disc 8–11 Gyr ago, and the Virgo Radial Merger (VRM), where the progenitor collided with the MW disc within the last 3 Gyr. These two scenarios make different predictions about observable structure in local phase space, because the morphology of debris depends on how long it has had to phase mix. The recently identified phase-space folds in Gaia DR3 have positive caustic velocities, making them fundamentally different than the phase-mixed chevrons found in simulations at late times. Roughly 20 per cent of the stars in the prograde local stellar halo are associated with the observed caustics. Based on a simple phase-mixing model, the observed number of caustics are consistent with a merger that occurred 1–2 Gyr ago. We also compare the observed phase-space distribution to FIRE-2 Latte simulations of GSE-like mergers, using a quantitative measurement of phase mixing (2D causticality). The observed local phase-space distribution best matches the simulated data 1–2 Gyr after collision, and certainly not later than 3 Gyr. This is further evidence that the progenitor of the ‘last major merger’ did not collide with the MW proto-disc at early times, as is thought for the GSE, but instead collided with the MW disc within the last few Gyr, consistent with the body of work surrounding the VRM.

milk waymergerstars
on-the-origin-of-the-moon-and-the-continents.pdf
on-the-origin-of-the-moon-and-the-continents.pdfon-the-origin-of-the-moon-and-the-continents.pdf
on-the-origin-of-the-moon-and-the-continents.pdf

All astronomical bodies originate inside clouds of gas and dust, therefore there should be a common process that leads to their condensation. In a galactic cloud there is always a gradient of speeds from point to point. Thanks to it, vortices originate that rake the material of the surrounding cloud gradually forming large gaseous disks, inside which vortices of second order develop that concentrate the matter of their orbits forming smaller and much denser disks, within which third order vortices further concentrate the matter. The dense cores of these vortices finally condense in massive bodies: sun, planets and satellites. The result should be a well ordered planetary system with no “debris” around and where both planets and satellites obey to a precise rule of the distances from their central body. The solar system complies with these conditions with three main exceptions. First, in an orbit where a large planet should be there is only a huge number of scattered asteroids. Second, Earth and its moon with all evidence were not formed in the same vortex, which means that Moon originated somewhere else. Third, Neptune’s satellite system has been shattered by the intrusion of a foreign body, Triton, and its largest satellites are missing. These exceptions seem to be strictly connected to each other and all due to a unique event, that is: Triton has diverted the largest Neptune satellite towards the Sun. This satellite impacted at high speed against the missing planet, scattering myriads of fragments from its mantle and pushing it towards the sun, where it eventually fell. The planet had at least 4 satellites some of which remained in their previous orbit, but two of them were dragged towards the sun and were captured by Earth. The largest became its lonely moon while the second fell on its surface giving origin to the continents. This event happened about 3,96 billion of years ago, as it is proven by the ages of the numerous samples brought from the moon

medwin publishersopen access journalopen access
p-value that measures the likelihood that the underlying
cumulative distribution was linear:
( ) ( )
( ) ( )
Q z e
p Q D n
2 1
, 10
j
j z j
1
1 2 2 2
å
= -
=
=
¥
- -
where D is the maximum distance between the data and linear
cumulative distributions and n is the number of data points. The
higher the p is, the more homogeneous the system is, and
therefore filament formation is unlikely, whereas a small p
indicates filament formation is more likely. Here we assume
that filament formation is very likely when p < 0.10, the same
as in previous work (Carrera et al. 2015).
For self-gravitating clump detection in our 3D simulations,
we use PLanetesimal ANalyzer (PLAN; Li et al. 2019). This is
a tool to identify self-gravitating clumps specifically made for
Athena output. The density of each particle is assessed based
on the nearest particles. Particles with densities higher than a
threshold are associated with neighboring dense particles until
a density peak is achieved. In contrast, particle groups with a
saddle point less than a threshold remain separated. Detailed
descriptions are found in previous work (Li et al. 2019).
2.3. Model Parameters
The main parameters for the Athena simulations are (1) the
dimensionless stopping time τf (the Newton regime; see
Section 1.2 and Nakajima et al. 2022), (2) the normalized
pressure parameter Δ = ηvK/cs, (3) the ratio of the particle surface
density to the gas surface density Z = Σp/Σg (VMF = Z
1
1 +
), and
(4) the normalized gravity G̃ G
4 g,0
2
p r
º W for 3D simulations,
where G is the gravitational constant. A small value of τf indicates
a small particle radius Rp, where the particle is well coupled with
the gas, whereas a large value of τf corresponds to a large value of
Rp, which is more decoupled from the gas. A large value of Δ
corresponds to a large pressure gradient and quicker radial infall.
G̃ represents the strength of the self-gravity. The parameter space
we are exploring for our 2D simulations is τf = 10−3
, 10−2
, 10−1
,
100
, 101
; Δ = 0.1, 0.2, 0.3, 0.4, 0.5; and Z = 0.05 and 0.1. For 3D
simulations, we use Z = 0.1 and 0.3 and G̃ 0.1788
= and 0.5898,
where the former G̃ value corresponds to a slightly cooler
temperature (4700 K), while the latter corresponds to a hotter
temperature (5200 K). Here, we justify the choice of these
parameters. This range of τf corresponds to the particle radii
between 2 m and 20 km (see Equation (5)). The global disk
structures have been calculated based on hydrodynamic simula-
tions in previous work. The overall disk mass of the Moon-
forming disk is typically a few percent of Earth, depending on the
impact model (MD/ML = 1.35–2.80, where MD is the disk mass
and ML is the lunar mass; Nakajima & Stevenson 2014). The
midplane disk temperature ranges from 3000 K to 7000 K, and the
radial range of the disk is ∼1–8 R⊕ (see Figure 5 in Nakajima &
Stevenson 2014). The disk temperature is ∼4000–5500 K at r = 3
R⊕. For the general rocky and icy impact-induced disks, the disk
temperature can vary, typically in the range of thousands of K for
vapor-rich disks (Nakajima et al. 2022). The pressure gradient can
vary, but the typical value of η is ∼0.02–0.06 based on impact
simulations (Nakajima et al. 2022). In the Moon-forming disk, the
value of Z increases as the disk cools (on a timescale of
10–102
yr). In other words, Z can initially be zero in an energetic
Moon-forming impact model (e.g., Lock et al. 2018) and
eventually becomes infinity as the disk materials condense and
the gas disappears. Thus, picking a value of Z means that we are
seeing physics at a specific time. Since we are primarily interested
in high-vapor disks (i.e., an early phase of the disk), we explore
the mass ratio range Z ä [0.05, 0.1] for our 2D simulations and
Z ä [0.1, 0.3] for our 3D simulations. We use the large value of
Z = 0.3 as a sensitivity test. Higher values of Z can be achieved as
the disk cools, but we focus on small values of Z („0.3) for
several reasons. First, at a larger value of Z (>0.3), the
conventional GI in the liquid part of the disk can form moonlets.
At Z 0.76
Z
1
1
=
+
, and the total (Σp + Σg) surface density at
r ∼ 3 R⊕ is ∼108
kg m−2
in energetic models, which means that
Σp = 0.76 × 108
kg m−2
= 7.6 × 107
kg m−2
. Previous work
suggests that when the liquid (melt) layer’s thickness reaches
5–10 km (or the equivalent surface density of a few 107
kg m−2
),
GI can happen in the melt layer (Machida & Abe 2004). Whether
streaming instability occurs at the same time or whether
streaming instability affects the GI in the Moon-forming disk
are unknown and have not been explored in previous studies.
Under these circumstances, the important of streaming instability
becomes unclear. Additionally, these streaming instability
simulations have been conducted at low-Z values („0.1) in
previous work to reproduce conditions in the protoplanetary
disks (e.g., Abod et al. 2019; Li & Youdin 2021), and
simulations with high-Z (>0.3) values have not been fully
tested. For these reasons, we focus on relatively small values of Z
in this study.
For the set of Athena simulations, we focus on reproducing
two sets of the Moon-forming disk thermal profile; assuming
M = M⊕ and r = 3 R⊕, where M⊕ is the Earth mass and R⊕ is
the Earth radius, the gas pressure at the midplane around 3 R⊕
is ≈12 MPa, T = 4700 K, and mm = 30 g mol−1
. This makes
the density ρg,0 = 12.13 kg m−3
for an ideal gas, cs = 1140 m
s−1
, vK = 4562 m s−1
, Ω = 2.38 × 10−4
s−1
, the scale height
H = cs/Ω = 4.78 × 106
m, and G̃ 0.1788
= . For the higher-
temperature scenario, T = 5200 K, ρg,0 = 40.01 kg m−3
,
cs = 1200 m s−1
, H = 5.03 × 106
, and G̃ 0.5898
= . These
temperatures are motivated by previous hydrodynamic calcula-
tions of the Moon-forming disk formation (Nakajima &
Stevenson 2014), and the other parameters are calculated
based on an equation of state of dunite assuming that the vapor
and liquid phases are in equilibrium (MANEOS; Thompson &
Lauson 1972; Melosh 2007). Since we are assuming that the
disk is in liquid–vapor equilibrium, a higher disk temperature
leads to a higher vapor density at the midplane. For example,
the gas density at liquid–vapor equilibrium is 150 kg m−3
at
6000 K and 1.8 kg m−3
at 4000 K for dunite, according to
MANEOS (Thompson & Lauson 1972; Melosh 2007). Assum-
ing η ∼ 0.02–0.06 for the Moon-forming disk, Δ ∼ 0.1–0.2.
However, higher values are possible (Nakajima et al. 2022);
therefore, we explore the range of Δ ∼ 0.1–0.5.
The parameter values of Δ and G̃ are significantly different
from values in the protoplanetary disk, where the typical values
used are Δ ∼ 10−3
–10−2
and G̃ 0.05
~ in the protoplanetary
disk (Carrera et al. 2015; Simon et al. 2016). We use a fixed
value of Z in hydrodynamic calculations because the
condensation timescale (∼years) is longer than the simulation
timescale (we run 2D simulations for ∼100 orbits, which
correspond to 123 hr. A steady state is reached by this time).
5
The Planetary Science Journal, 5:145 (12pp), 2024 June Nakajima et al.
We also assume that the disk does not evolve on this short
timescale.
3. Results
3.1. 2D Athena Simulations
Figure 1(A) shows four examples of spacetime plots of our
2D simulations, where (1) (τf, Δ) = 10−3
, 0.3, (2) 10−2
, 0.1,
(3) 10−1
, 0.2, and (4) 10−1
, 0.2. Z = 0.1 for the four cases.
These simulations represent four characteristic regimes. The
horizontal axis is x normalized by the scale height H, and the
vertical axis is the number of orbits. The color shows particle
concentration. In case 1, τf is small, which means that gas and
particles are well coupled, and this does not lead to filament
formation. In case 2, after ∼20 orbits, filaments start forming,
and these filaments are stable during the rest of the simulations,
which indicate that streaming instability occurs with this
chosen parameter. A radially shearing periodic boundary
condition is used in the x-direction; therefore, the filament that
appears at x/H = −0.1 at ∼100 orbits reappears at x/H = 0.1.
Two distinct filaments form in this simulation. In case 3,
filaments form, but they are not stable because their radial
movements are high due to the high Δ value. Thus, this is not
an ideal parameter space for streaming instability. Similar
behaviors have been seen in simulations for the protoplanetary
disk with other parameter combinations (e.g., Δ = 0.05, τf = 3,
Z = 0.005; Carrera et al. 2015). In case 4, the filaments are not
as clearly defined as those in case 2, but a coherent filament
structure is found after several orbits. Thus, this is also
considered as a streaming instability regime.
Figure 1(B) shows the summary of our 2D simulations for
Z = 0.1 (top) and Z = 0.05 (bottom). We use the KS test to
identify the extent of particle concentration (see Section 2.2).
Concentration is measured by the p-value, and when the value of
p is small (p < 0.1), streaming instability is likely (Carrera et al.
2015). The color bar indicates the p-value. For Z = 0.1, p < 0.1
is achieved when Δ = 0.1 and 10−3
„ τf „ 100
. For large Δ
values, the radial motions are large, and stable filaments do not
form, as discussed above. For Z = 0.05, this trend remains the
same, but the streaming instability regime is smaller
(10−2
„ τf „ 100
) than that for Z = 0.1. This is because the
larger Z leads to the presence of more particles in the disk and
therefore to more filament formation (Carrera et al. 2015). Thus,
our 2D simulations show that the most filaments form at
relatively small Δ (Δ = 0.1) and with both Z = 0.1 and Z = 0.05,
but the higher Z has more favorable conditions. This general
trend is consistent with previous work in the protoplanetary disk
(e.g., Carrera et al. 2015), which investigates smaller Δ values
for the protoplanetary disk (Δ „ 0.05).
3.2. 3D Athena Simulations
Now that we have identified the parameter space for the
streaming instability (Δ = 0.1, Z = 0.1), we perform 3D
simulations with self-gravity to estimate the size distributions
of streaming-instability-induced moonlets. Figure 2 shows
Figure 1. (A) Spacetime diagram for the four cases. The input parameters are (1) (τf, Δ) = 10−3
, 0.3, (2) 10−2
, 0.1, (3) 10−1
, 0.2, and (4) 1, 0.1, all at Z = 0.1. The
horizontal axis is x normalized by the scale height H. The vertical axis indicates the number of orbits. The color shows p p
S áS ñ, where Σp is the particle surface
density and p
áS ñ is the average along the x-axis. Filament formation occurs in cases 2 and 4, while no filament formation occurs in cases 1 and 3. (B) Summary of
results for Z = 0.1 and Z = 0.05. The colors show the p-value, and the clumping regime (p < 0.1) is shown in the sky blue shade. Parameters for cases 1–4 are
indicated. This shows that clumping occurs only at small Δ values (Δ = 0.1).
6
The Planetary Science Journal, 5:145 (12pp), 2024 June Nakajima et al.
snapshots of one of our 3D simulations ( G̃
1,
f
t = =
Z
0.5898, 0.1
= , and Δ = 0.1) at four different times
(t = 0.32, 2.87, 3.39, and 3.55), where t is the number of
orbits. The self-gravity is not initially included and is turned on
at t = 3.18. This is a general practice to avoid artificial
clumping before streaming instability takes place (Simon et al.
2016). The color shows the particle density. The circles
indicate the Hill spheres of self-bound clumps, which are
identified using PLAN (Li et al. 2019). At t = 0.32, no clump is
identified, but by t = 2.87, the streaming instability fully
develops and reaches its steady state. After self-gravity is
turned on at t = 3.18, self-gravitating clumps form right away
(t = 3.39). This general behavior is similar to previous work on
the streaming instability in the protoplanetary disk (Abod et al.
2019), but the time it takes to form clumps by streaming
instability is shorter, probably because of the large Z value
(Simon et al. 2022) compared to those of the protoplane-
tary disk.
Figure 3(A) shows the cumulative distribution of moonlet
mass Mp, normalized by the characteristic self-gravitating mass
MG, defined as (Abod et al. 2019)
⎛
⎝
⎞
⎠
˜ ( ) ( )
M
G
Z G H
2
4
2
2
, 11
G
G
p
p
2
5
2 3
4
9 2 3 2
g,0
3
p
l
p
p r
= S =
S
W
=
where λG is an instability wavelength, which originates from
the Toomre dispersion relation, equating the tidal and
gravitational forces (Abod et al. 2019). The parameters are
τf = 0.1, 1, G̃ 0.1788, 0.5898
= , and Z = 0.1, 0.3. For the
G̃ 0.1787
= cases, MG = 5.19 × 1018
kg, and for the
G̃ 0.5898
= cases, MG = 2.17 × 1020
kg. For all 3D simula-
tions, Δ = 0.1 is assumed. The solid lines indicate the moonlet
mass distribution shortly after the onset of the self-gravity
(t = 3.66 for G̃ 0.1788, 0.1
f
t
= = , and t = 3.34 for all the
other cases). The dashed–dotted lines indicate the same
parameter after an additional time (t + Δt, where Δt = 1/
Ω = 0.16 orbits, except the Z = 0.3 case, where t
1
2
D = W
). The
Figure 2. Snapshots of a 3D simulation (τf = 1, Δ = 0.1, Z = 0.1, G̃ 0.5898
= ). The horizontal and vertical axes are x and y, normalized by the scale height H. The
color indicates the particle density, normalized by the average density. t indicates the number of orbits. The circles represent the Hill radius of each moonlet formed by
streaming instability. At first (t = 0.32), no concentration of particles is observed, but streaming instability clearly develops by t = 2.87. After self-gravity is turned on
at t = 3.18, moonlets form by GI (t = 3.39, 3.55). The self-gravitating clumps are identified using PLAN (see the main text).
7
The Planetary Science Journal, 5:145 (12pp), 2024 June Nakajima et al.
reason for the shorter Δt for the higher-Z case is that the
streaming instability develops faster for higher-Z values (Simon
et al. 2022). The maximum Mp/MG = 0.254 (G̃ 0.1788,
=
1
f
t = at t = 3.50). This is broadly consistent with previous
studies that suggest that the maximum clump masses formed by
the streaming instability are characterized by MG in the range of
10−1
−101
MG (Abod et al. 2019; Li et al. 2019). Our result lies
on the lower end of this mass range. This indicates that the
moonlet mass distribution based on our numerical simulations
is consistent with the analytical mass model generated for the
protoplanetary disk. Thus, this also indicates that the processes
of streaming instability are similar between the Moon-forming
disk and protoplanetary disk despite the different input
parameters. We also find that higher Z does not necessarily
lead to higher Mp.
Here we consider the best-case scenario of forming a large
moonlet. Assuming that the density of the clump is 3000 kg m−3
,
MG = 2.17 × 1020
kg is equivalent to 258 km in radius and 0.254
MG = 163 km. Based on Equation (1), the residence times of
258 km and 163 km moonlets are 92 and 58 days, respectively,
assuming vr,g = 0. These timescales are much shorter than the
Moon-formation timescale (10 s–100 yr, Thompson & Steven-
son 1988). Thus, even though the streaming instability can occur
in an initially vapor-rich Moon-forming disks, it does not help
increase the residence time of moonlets.
3.3. Exomoon Formation by Streaming Instability
The streaming instability likely occurs in impact-induced
moon-forming disks in extrasolar systems. The pressure gradient
parameter η (∼0.02–0.06) is similar regardless of the composition
of the disk (Nakajima et al. 2022). Figure 3(B) shows the disk
residence time for a clump with mass MG in a moon-forming disk
as a function of the planetary mass (Mplanet = 1–6 M⊕). “Rocky
planet” corresponds to disks formed by a collision between
terrestrial planets, while “icy planet” indicates disks formed by a
collision between icy planets whose mantles are made of water ice
(70 wt%) and cores are made of iron (30 wt%). This is produced
by calculating r/vr (see Equation (1)), where for rocky planets, we
use ρg,0 = 40.01 kg m−3
, T = 5200 K, ρp = 3000 kg m−3
, and
r = 3 Rplanet, where Rplanet is the planetary radius, and for icy
planets we use ρg,0 = 10.0 kg m−3
, T = 2000 K, ρp =
1000 kg m−3
, and r = 3 Rplanet. These are the temperatures when
the impact-induced disks reach complete vaporization based on
impact simulations (Nakajima et al. 2022). A higher temperature
is possible, but the disk would need to cool down to this
temperature to form particles (dust); therefore, this is the most
relevant temperature to assess the effect of streaming instability.
The thermal state of the disk formed by icy planetary collisions is
estimated based on an equation of state of water (Senft &
Stewart 2008). Additionally, ( )
R R M M
planet planet
1 3.7
= Å Å for
rocky planets and ( )
R R M M
1.2
planet planet
1 3
= Å Å for icy planets
are assumed (Kipping et al. 2013; Mordasini et al. 2012). At
Mplanet = 1–6 M⊕, these parameters make G̃ 0.42 0.67
= - for
rocky planets and G̃ 0.25
= for icy planets, which are similar to
the ranges covered in our hydrodynamic simulations (Section 2.3).
In both icy and rocky planet cases, the disk residence time is
a few tens of days to several months, which is short compared
to the satellite formation timescale (10 s–100 s yr, Nakajima
et al. 2022). This suggests that the streaming instability likely
plays a limited role in impact-induced moon-forming disks.
4. Discussion
4.1. Streaming Instability in the Moon-forming Disk
Figure 4 shows a schematic view of our hypothesis. An
energetic impact would generate a vapor-rich disk (the disk is
made of silicate vapor for the Moon-forming disk and rocky
moon-forming disks and water vapor for icy moon-forming
disks). Over time, the disk cools by radiation and small droplets
(<cm) emerge. These small droplets would grow by accretion
and streaming instability. However, once these moonlets reach
100 m–100 km in size, gas drag from the vapor is so strong that
they fall onto the planet on a short timescale (days–weeks).
This continues to occur until the VMF of the disk decreases so
that the gas drag effect is no longer strong (we assume the
silicate vapor is in equilibrium with the silicate liquid. As the
disk temperature decreases, the gas density decreases). Once
this condition is reached, a liquid layer emerges, and moonlets
can stay in the disk. However, by this time, a significant disk
mass could have been lost (>80 wt%; Ida et al. 2020; Nakajima
et al. 2022). For this reason, only a small moon (or moons)
Figure 3. (A) Cumulative mass distribution of moonlets formed by streaming instability for Δ = 0.1, Z = 0.1, 0.3. The purple, dark blue, light blue, dark yellow, and
light yellow lines indicate parameter values of ( ˜ )
G Z
, ,
f
t = (0.5898, 0.1, 0.1), (0.1788, 0.1, 0.1), (0.5898, 1, 0.1), (0.1788, 1, 0.1), and (0.1788, 0.1, 0.3). The solid and
dashed–dotted lines represent the mass distribution at different times (see Section 3.2). The horizontal axis indicates the moonlet mass (Mp) normalized by MG, and the
vertical axis indicates the number of moonlets whose masses are larger than the given moonlet mass. (B) Residence time of a moonlet whose mass is MG formed by
streaming instability at r = 3 Rplanet as a function of the planetary mass Mplanet normalized by the Earth mass. The blue solid and yellow dashed–dotted lines represent
disks formed by collisions between icy planets and rocky planets, respectively.
8
The Planetary Science Journal, 5:145 (12pp), 2024 June Nakajima et al.

Recommended for you

A Tale of 3 Dwarf Planets: Ices and Organics on Sedna, Gonggong, and Quaoar f...
A Tale of 3 Dwarf Planets: Ices and Organics on Sedna, Gonggong, and Quaoar f...A Tale of 3 Dwarf Planets: Ices and Organics on Sedna, Gonggong, and Quaoar f...
A Tale of 3 Dwarf Planets: Ices and Organics on Sedna, Gonggong, and Quaoar f...

The dwarf planets Sedna, Gonggong, and Quaoar are interesting in being somewhat smaller than the methane-rich bodies of the Kuiper Belt (Pluto, Eris, Makemake), yet large enough to be spherical and to have possibly undergone interior melting and differentiation. They also reside on very different orbits, making them an ideal suite of bodies for untangling effects of size and orbit on present day surface composition. We observed Sedna, Gonggong, and Quaoar with the NIRSpec instrument on the James Webb Space Telescope (JWST). All three bodies were observed in the low-resolution prism mode at wavelengths spanning 0.7 to 5.2 μm. Quaoar was additionally observed at 10x higher spectral resolution from 0.97 to 3.16 μm using mediumresolution gratings. Sedna’s spectrum shows a large number of absorption features due to ethane (C2H6), as well as acetylene (C2H2), ethylene (C2H4), H2O, and possibly minor CO2. Gonggong’s spectrum also shows several, but fewer and weaker, ethane features, along with stronger and cleaner H2O features and CO2 complexed with other molecules. Quaoar’s prism spectrum shows even fewer and weaker ethane features, the deepest and cleanest H2O features, a feature at 3.2 μm possibly due to HCN, and CO2 ice. The higher-resolution medium grating spectrum of Quaoar reveals several overtone and combination bands of ethane and methane (CH4). Spectra of all three objects show steep red spectral slopes and strong, broad absorptions between 2.7 and 3.6 μm indicative of complex organic molecules. The suite of light hydrocarbons and complex organic molecules are interpreted as the products of irradiation of methane. The differences in apparent abundances of irradiation products among these three similarly-sized bodies are likely due to their distinctive orbits, which lead to different timescales of methane retention and to different charged particle irradiation environments. In all cases, however, the continued presence of light hydrocarbons implies a resupply of methane to the 2 surface. We suggest that these three bodies have undergone internal melting and geochemical evolution similar to the larger dwarf planets and distinct from all smaller KBOs. The feature identification presented in this paper is the first step of analysis, and additional insight into the relative abundances and mixing states of materials on these surfaces will come from future spectral modeling of these data.

james webbquaoarsedna
Younger moon
Younger moonYounger moon
Younger moon

1) New measurements of tungsten isotopes in lunar rocks indicate that the Moon formed later than previously thought, between 62-150 million years after the formation of the solar system, challenging current models of early planetary formation. 2) This later formation of the Moon requires revising our understanding of the timing of events like the giant impact that formed the Earth-Moon system and the solidification of the lunar magma ocean. 3) The new timeline suggests Earth's core may have formed independently of the giant impact and that magma oceans on Earth and other terrestrial planets took longer to solidify than models predicted.

Tidal star-planet interaction and its observed impact on stellar activity in ...
Tidal star-planet interaction and its observed impact on stellar activity in ...Tidal star-planet interaction and its observed impact on stellar activity in ...
Tidal star-planet interaction and its observed impact on stellar activity in ...

This study investigates tidal star-planet interaction by analyzing X-ray observations of planet-hosting stars that are part of wide binary systems. The study compares the X-ray luminosity, an indicator of stellar activity, between the planet-hosting primary star and its non-planet hosting companion star. By analyzing 34 such binary systems with X-ray data from XMM-Newton and Chandra, the study finds that planet hosts with close-in massive planets have higher X-ray luminosity compared to their companion stars, indicating tidal interaction is altering the rotation and magnetic activity of the host stars. The study concludes tidal interaction from massive, close-in planets can impact the rotational evolution of stars, while more distant or smaller planets likely

júpiter quenteexoplaneta
could form from an initially vapor-rich disk. Thus, we suggest
that an initially vapor-rich disk is not suitable for forming our
Moon, and our result supports the hypothesis that the Moon
formed from an initially vapor-poor disk, including the
canonical model where the proto-Earth was hit by a Mars-
sized impactor (Canup & Asphaug 2001).
The isotopic composition of the Moon may constrain whether
streaming instability played a role during the Moon formation.
Some may argue that the observational fact that the Moon is
depleted in volatiles may be explained by accreting moonlets
formed by streaming instability before volatiles accreted onto the
Moon. However, this process would make the Moon isotopically
light if these isotopes experience kinetic fractionation (Dauphas
et al. 2015, 2022), which is inconsistent with the observation of
the enrichment of heavy isotopes in some elements in the Moon
(such as K; Wang & Jacobsen 2016). The lunar isotopes would be
heavier if they experience equilibrium fractionation, but the
equilibrium fractionation would not produce the observed isotopic
fractionations under the high disk temperature condition. Alter-
natively, some volatiles could be lost from the lunar magma ocean
under an equilibrium condition (Charnoz et al. 2021), but the
efficiency of the volatile loss is not fully known (Dauphas et al.
2022).
4.2. Streaming Instability in General Impact-induced Disks
Our model supports the previous work that suggests that
relatively large rocky (>6 M⊕, >1.6 R⊕) and icy (>1 M⊕, >
1.3 R⊕) planets cannot form impact-induced moons that are
large compared to the host planets (Nakajima et al. 2022).
Larger planets than those thresholds generate completely vapor
disks, because the kinetic energy involved in an impact scales
with the planetary mass. Thus, these large planets are not
capable of forming moons that are large compared to their host
planets. Moons can form by mechanisms other than impacts,
such as formation in a circumplanetary disk and gravitational
capture, but these moons tend to be small compared to the sizes
of their host planets (the predicted moon-to-planet mass ratio is
∼10−4
for moons formed by circumplanetary disk, Canup &
Ward 2006; and gravitationally captured moons are small in the
solar system, Agnor & Hamilton 2006). Thus, fractionally large
moons compared to the host planet sizes, which are
observationally favorable, likely form by impact. So far, no
exomoon has been confirmed despite extensive searches, but
future observations, especially with the James Webb Space
Telescope (Cassese & Kipping 2022), may be able to find
exomoons and test this theoretical hypothesis.
4.3. Comparison between the Moon-forming Disk and
Protoplanetary Disk
It is certainly intriguing that the streaming instability is able to
solve the gas drag problem in the protoplanetary disk but not in
the Moon-forming disk. In both scenarios, the clump sizes
formed by the streaming instability happen to be similar
(MG ∼ 100 km, see Johansen et al. 2015, for the protoplanetary
disk and ∼100 km for the protolunar disk, respectively). This
size is ∼105
times larger than the size of the particle (0.52–1 m;
Adachi et al. 1976) that experiences the strongest gas drag
(τf = 1 and vr ∼ −ηvK) in the protoplanetary disk. Therefore,
once particles become ∼100 km-sized planetesimals by
streaming instability, the radial velocities of the planetesimals
decrease drastically (vr
v
2 K
f
= -
h
t
for large τf; see Equation (A7)),
which helps planetesimal growth. In contrast, the largest moonlet
size (100 km) formed by streaming instability in the Moon-
forming disk is only 50 times larger than the particle size that
experiences the strongest gas drag (2 km). This does not result in
a large τf change or vr change; therefore, streaming instability
does not effectively help moon formation. Therefore, streaming
instability is an effective mechanism to skip the “1 m barrier” in
Figure 4. Schematic view of Moon formation from an initially vapor-rich disk.
9
The Planetary Science Journal, 5:145 (12pp), 2024 June Nakajima et al.
the protoplanetary disk, whereas it is not an effective mechanism
to skip the “kilometer barrier” in the Moon-forming disk.
4.4. Streaming Instability in the Circumplanetary Disk around
a Gas Giant
The streaming instability may occur during moon formation in
various satellite systems around gas giants. The Galilean moons
around Jupiter and Titan around Saturn likely formed from their
circumplanetary disks. In these disks, the particle-to-gas ratio Z is
typically 10−4
–10−2
, which was thought to be too small for
streaming instability to occur (Carrera et al. 2015; Yang et al.
2017; Shibaike et al. 2017). However, recent streaming instability
calculations show that streaming instability can occur as low as
Z = 4 × 10−3
(at Δ = 0.05 and τf = 0.3; Li & Youdin 2021), and
the required Z depends on the disk conditions (e.g., Carrera et al.
2015; Yang et al. 2017; Sekiya & Onishi 2018). This may
indicate that streaming instability can be potentially important in
circumplanetary disks around gas giants if the dust–gas ratio of
the disk is relatively high. The velocity difference between the gas
and particles vrel is (Canup & Ward 2002)
⎛
⎝
⎞
⎠
( )
v v c
c
r
. 12
rel k s
s
h
= ~
W
Assuming 0.1
c
r
c
vk
s s
= ~
W
(Canup & Ward 2002), this yields
η ∼ 0.01 and 0.1
v
c
k
s
D = ~
h
. If Z is sufficiently large (at least
Z > 4 × 10−3
; Li & Youdin 2021), it is possible that the
streaming instability takes place in a circumplanetary disk
around a gas giant. A rough estimate of the size of a moonlet in
a circumplanetary disk is MG = 1.4 × 1016
kg (Equation (11)),
which is ∼10.3 km in radius if it is a rocky moonlet. Here, we
are assuming Z = 10−2
, Σp = 3 × 104
kg m−2
, cs = 1000 m s−1
,
H ∼ cs/Ω, and ( )
H 2
g g
r p
= S (Canup & Ward 2002). These
moonlets are relatively small compared to large moons around
gas giants (e.g., the Galilean moon masses are in the range of
1022
–1023
kg), and it is unclear if they would significantly impact
these moon-formation processes. Further research is needed to
understand their effects on the satellite formation in a
circumplanetary disk around a gas giant.
4.5. Model Limitations
There are several model limitations that need to be addressed in
our future work. First, the effect of the Roche radius (aR ∼ 3 R⊕)
needs to be taken into account to understand the Moon's
formation. An inspiraling moonlet would not directly reach the
Earth but would be tidally disrupted near the Roche radius, where
it then might be incorporated into an accretion disk (e.g., Salmon
& Canup 2014). Second, our model presented here does not take
into account the evolution of the disk in detail, which is important
to identify the final mass and composition of the resulting moon.
As the disk spreads out, it is likely that Δ decreases, which would
slow down the radial drift of moonlets. This gas drag effect would
disappear once the local vapor condenses, which would occur at
the outer part of the disk first, since that part of the disk can cool
efficiently due to its large surface area. Given that such moonlets
that directly form from the disk would not have time to lose
volatiles in the disk phase, this model requires the Moon to have
lost its volatiles before or after the disk phase, for example, during
the lunar magma ocean phase (Charnoz et al. 2021; see further
discussion in Section 4.1). Nevertheless, such a scenario may be
possible in moon-forming disks in the solar and extrasolar
systems. The effects of the Roche radius and disk evolution will
be addressed in our future work.
5. Conclusions
In conclusion, we conduct hydrodynamic simulations using
Athena and show that the streaming instability can form self-
gravitating clumps (∼102
km) in a vapor-rich moon-forming disk
generated by a giant impact, but the sizes of the clumps it
generates are not large enough to avoid inspiraling due to the
strong gas drag. This is a major difference from processes in the
protoplanetary disk, where the streaming instability can efficiently
form large clumps to avoid the strong gas drag effect. As a result,
growing moonlets in an initially vapor-rich moon-forming disk
continue to fall onto the planet once they reach sizes of 100 m–
100 km. These moonlets could grow further once the disk cools
enough and the VMF of the disk becomes small. However, by this
time, a significant amount of the disk mass is lost, and the
remaining disk could make only a small moon. This result is
applicable to impact-induced moon-forming disks in the solar
system and beyond; we find that the streaming instability is not an
efficient mechanism to form a large moon from an impact-induced
vapor-rich disk in general. As a result, we support previous work
that suggests that fractionally large moons compared to their host
planets form from vapor-poor disks. This supports the Moon-
formation models that produce vapor-poor disks, such as the
canonical model. For exomoons, our work supports the previous
work that suggests that the ideal planetary radii that host
fractionally large moons are 1.3–1.6 R⊕ (Nakajima et al. 2022)
given that rocky or icy planets larger than these sizes would likely
produce completely vapor disks, which are not capable of forming
large moons (Nakajima et al. 2022). The streaming instability may
take place in circumplanetary disks, but their effect on the moon-
formation process needs further investigation.
Acknowledgments
We thank the code developers of Athena (Stone et al. 2008;
Bai & Stone 2010a; Simon et al. 2016) and PLAN (Li et al.
2019). We appreciate discussions with Shigeru Ida and Scott D.
Hull. J.A. was partially supported by the Research Experience
for Undergraduates (REU) Program, National Science Founda-
tion (NSF), under grant No. PHY-1757062. M.N. was supported
in part by the National Aeronautics and Space Administration
(NASA) grant Nos. 80NSSC19K0514, 80NSSC21K1184, and
80NSSC22K0107. Partial funding for M.N. was also provided
by NSF EAR-2237730 as well as the Center for Matter at
Atomic Pressures (CMAP), an NSF Physics Frontier Center,
under award PHY-2020249. Any opinions, findings, conclu-
sions, or recommendations expressed in this material are those of
the authors and do not necessarily reflect those of the National
Science Foundation. M.N. was also supported in part by the
Alfred P. Sloan Foundation under grant G202114194.
Appendix
Appendix A
The momentum equation for gas in the radial direction is
written as
( )
v
r
GM
r
dp
dr
1
, A1
2
2
g
g
r
= +
f Å
10
The Planetary Science Journal, 5:145 (12pp), 2024 June Nakajima et al.
where vf is the azimuthal velocity of the gas. Using
Equation (2) in the main text and the relationship of p c
g g 2
2
r
= ,
( ) ( )
v v 1 2 . A2
K
1
2
h
= -
f
The equations of motion of the particles in the gas in the radial
and azimuthal directions, vr and vf, are (Armitage 2010;
Takeuchi & Lin 2002)
( ) ( )
dv
dt
v
r
r
t
v v
1
, A3
r
r r
2
2
f
,g
= - W - -
f
( ) ( ) ( )
d
dt
rv
r
t
v v , A4
f
,g
= - -
f f f
where vr,g and vf,g are the gas velocities in the radial and
azimuthal directions. Assuming that the
d
dt
terms are negligible,
one finds that
( )
v
v v
2
. A5
r
r
f
1
,g K
f f
1
t h
t t
=
-
+
-
-
This formulation is slightly different from previous formula-
tions (−2ηvK instead of −ηvK; Armitage 2010; Takeuchi &
Lin 2002) due to the different definition of η by a factor of 2.
This leads to
( )
v
v v
2
. A6
r
r
f
1
,g K
f f
1
t h
t t
=
-
+
-
-
This can be further simplified by the values of τf as
⎧
⎨
⎪
⎪
⎩
⎪
⎪


( )
( )
v
v v
v v
v
2 at 1,
1
2
2 at 1,
2
at 1.
A7
r
r
r
K
,g f K f
,g K f
f
f
ht t
h t
h
t
t
=
-
- ~
-
For a steady disk flow, one could assume vr r
,g
3
2
~ -
n
(Armitage 2010). In the main text, we simply assume vr,g = 0.
Appendix B
An example input file for a 2D Athena simulation is listed in
the supplementary material (athinput.par_strat2d). The resolu-
tion in this file is 512 × 512, and τf = 1, Z = 0.1, and Δ = 0.1.
ORCID iDs
Miki Nakajima https:/
/orcid.org/0000-0001-5014-0448
Jeremy Atkins https:/
/orcid.org/0009-0005-0722-8408
Jacob B. Simon https:/
/orcid.org/0000-0002-3771-8054
Alice C. Quillen https:/
/orcid.org/0000-0003-1280-2054
References
Abod, C. P., Simon, J. B., Li, R., et al. 2019, AJ, 883, 192
Adachi, I., Hayashi, C., & Nakazawa, K. 1976, PThPh, 56, 1756
Agnor, C. B., & Hamilton, D. P. 2006, Natur, 441, 192
Armitage, P. J. 2010, Astrophysics of Planet Formation (Cambridge:
Cambridge Univ. Press)
Armytage, R. M. G., Georg, R. B., Williams, H. M., & Halliday, A. N. 2012,
GeCoA, 77, 504
Asphaug, E., Emsenhuber, A., Cambioni, S., Gabriel, T. S. J., &
Schwartz, S. R. 2021, PSJ, 2, 200
Bai, X.-N., & Stone, J. M. 2010a, ApJ, 722, 1437
Bai, X.-N., & Stone, J. M. 2010b, ApJS, 190, 297
Bonomo, A. S., Zeng, L., Damasso, M., et al. 2019, NatAs, 3, 416
Bouhifd, A. M., Jephcoat, A. P., Porcelli, D., Kelley, S. P., & Marty, B. 2020,
GChPL, 15
Cameron, A. G. W., & Ward, W. R. 1976, Lunar Planet. Sci. VII, 120
Cano, E. J., Sharp, Z. D., & Shearer, C. K. 2020, NatGe, 13, 270
Canup, R. M. 2004, Icar, 168, 433
Canup, R. M. 2005, Sci, 307, 546
Canup, R. M. 2012, Sci, 338, 1052
Canup, R. M., & Asphaug, E. 2001, Natur, 412, 708
Canup, R. M., Visscher, C., Salmon, J., & Fegley, B. 2015, NatGe, 8, 918
Canup, R. M., & Ward, W. R. 2002, AJ, 124, 3404
Canup, R. M., & Ward, W. R. 2006, Natur, 441, 834
Canup, R. M., Righter, K., Dauphas, N., et al. 2023, RvMG, 89, 53
Carrera, D., Johansen, A., & Davies, M. B. 2015, A&A, 579, A43
Cassese, B., & Kipping, D. 2022, MNRAS, 516, 3701
Charnoz, S., & Michaut, C. 2015, Icar, 260, 440
Charnoz, S., Sossi, P. A., Lee, Y.-N., et al. 2021, Icar, 364, 114451
Chiang, E., & Youdin, A. N. 2010, AREPS, 38, 493
Colella, P. 1990, JCoPh, 87, 171
Colella, P., & Woodward, P. R. 1984, JCoPh, 54, 174
Craddock, R. A. 2011, Icar, 211, 1150
Cúk, M., Hamilton, D. P., Lock, S. J., & Stewart, S. T. 2016, Natur, 539, 402
Ćuk, M., Lock, S. J., Stewart, S. T., & Hamilton, D. P. 2021, PSJ, 2, 147
Cúk, M., & Stewart, S. T. 2012, Sci, 338, 1047
Dauphas, N. 2017, Natur, 541, 521
Dauphas, N., Poitrasson, F., Burkhardt, C., Kobayashi, H., & Kurosawa, K.
2015, E&PSL, 427, 236
Dauphas, N., Nie, N. X., Blanchard, M., et al. 2022, PSJ, 3, 29
Halliday, A. N., & Canup, R. M. 2022, NRvEE, 4, 19
Halliday, N. H. 2004, Natur, 427, 505
Hartmann, W. K., & Davis, D. R. 1975, Icar, 24, 504
Hosono, N., Karato, S.-i., Makino, J., & Saitoh, T. R. 2019, NatGe, 12, 418
Hull, S. D., Nakajima, M., Hosono, N., Canup, R. M., & Gassmöller, R. 2024,
PSJ, 5, 9
Ida, S., Ueta, S., Sasaki, T., & Ishizawa, Y. 2020, NatAs, 4, 880
Johansen, A., Mac Low, M. M., Lacerda, P., & Bizzarro, M. 2015, SciA, 1,
e1500109
Johansen, A., Oishi, J. S., Mac Low, M. M., et al. 2007, Natur, 448, 1022
Kegerreis, J. A., Ruiz-Bonilla, S., Eke, V. R., et al. 2022, ApJL, 937, L40
Kenworthy, M., Lock, S., Kennedy, G., et al. 2023, Natur, 622, 251
Kipping, D. M., Forgan, D., Hartman, J., et al. 2013, ApJ, 777, 134
Krapp, L., Benítez-Llambay, P., Gressel, O., & Pessah, M. E. 2019, ApJL,
878, L30
Kruijer, T. S., Archer, G. J., & Kleine, T. 2021, NatGe, 14, 714
Kruijer, T. S., Kleine, T., Fischer-Godde, M., & Sprung, P. 2015, Natur,
520, 534
Li, R., & Youdin, A. N. 2021, ApJ, 919, 107
Li, R., Youdin, A. N., & Simon, J. B. 2019, ApJ, 885, 69
Lock, S. J., Stewart, S. T., Petaev, M. I., et al. 2018, JGRE, 123, 910
Machida, R., & Abe, Y. 2004, ApJ, 617, 633
Melosh, H. J. 2007, M&PS, 42, 2079
Meng, H. Y. A., Su, K. Y. L., Rieke, G. H., et al. 2014, Sci, 345, 1032
Mordasini, C., Alibert, Y., Georgy, C., et al. 2012, A&A, 547, 112
Mullen, P. D., & Gammie, C. F. 2020, ApJL, 903, L15
Nakajima, M., Genda, H., Asphaug, E., & Ida, S. 2022, NatCo, 13, 568
Nakajima, M., & Stevenson, D. J. 2014, Icar, 233, 259
Nakajima, M., & Stevenson, D. J. 2015, E&PSL, 427, 286
Nakajima, M., & Stevenson, D. J. 2018, E&PSL, 487, 117
Nie, N. X., & Dauphas, N. 2019, ApJL, 884, L48
Nielsen, S. G., Bekaert, D. V., & Auro, M. 2021, NatCo, 12, 1817
Pahlevan, K., & Stevenson, D. J. 2007, E&PSL, 262, 438
Rufu, R., Aharonson, O., & Perets, H. B. 2017, NatGe, 10, 89
Rufu, R., & Canup, R. M. 2020, JGRE, 125, e2019JE006312
Salmon, J., & Canup, R. M. 2012, ApJ, 760, 83
Salmon, J., & Canup, R. M. 2014, RSPTA, 372, 20130256
Sekiya, M., & Onishi, I. K. 2018, ApJ, 860, 140
Senft, L. E., & Stewart, S. T. 2008, M&PS, 43, 1993
Shibaike, Y., Okuzumi, S., Sasaki, T., & Ida, S. 2017, ApJ, 846, 81
Simon, J. B., Armitage, P. J., Li, R., & Youdin, A. N. 2016, ApJ, 822, 55
Simon, J. B., Blum, J., Birnstiel, T., & Nesvorný, D. 2022, Comets III, in press
(arXiv:2212.04509)
Slattery, W. L., Benz, W., & Cameron, A. G. W. 1992, Icar, 99, 167
Sossi, P. A., Moynier, F., & van Zuilen, K. 2018, PNAS, 115, 10920
Stone, J. M., & Gardiner, T. A. 2010, ApJS, 189, 142
Stone, J. M., Gardiner, T. A., Teuben, P., Hawley, J. F., & Simon, J. B. 2008,
ApJS, 178, 137
11
The Planetary Science Journal, 5:145 (12pp), 2024 June Nakajima et al.
Takahashi, S. Z., & Inutsuka, S.-i. 2014, ApJ, 794, 55
Takeuchi, T., & Lin, D. N. C. 2002, ApJ, 581, 1344
Thiemens, M. M., Sprung, P., Fonseca, R. O. C., Leitzke, F. P., & Münker, C.
2019, NatGe, 12, 696
Thiemens, M. M., Tusch, J., O. C. Fonseca, R., et al. 2021, NatGe, 14,
716
Thompson, C., & Stevenson, D. J. 1988, ApJ, 333, 452
Thompson, M. A., Weinberger, A. J., Keller, L. D., Arnold, J. A., &
Stark, C. C. 2019, ApJ, 875, 45
Thompson, S. L., & Lauson, H. S. 1972, Improvements in the Chart D
Radiation-Hydrodynamic Code. III: Revised Analytic Equations of State
Tech. Rep. SC-RR-71-0714, Sandia National Laboratory
Tominaga, R. T., Inutsuka, S.-i., & Takahashi, S. Z. 2018, PASJ, 70, 3
Tominaga, R. T., Takahashi, S. Z., & Inutsuka, S.-I. 2019, ApJ, 881, 53
Toro, E. F. 1999, Riemann Solvers and Numerical Methods for Fluid
Dynamics: A Practical Introduction (Dordrecht: Springer)
Touboul, M., Puchtel, I. S., & Walker, R. J. 2015, Natur, 520, 530
Wang, K., & Jacobsen, S. B. 2016, Natur, 538, 487
Ward, W. R., Canup, R. M., & Rufu, R. 2020, JGR, 125, e2019JE00626
Weidenschilling, S. J. 1977, MNRAS, 180, 57
Wiechert, U., Halliday, A. N., Lee, D. C., et al. 2001, Sci, 294, 345
Williams, C. D., & Mukhopadhyay, S. 2019, Natur, 565, 78
Williams, C. D., Mukhopadhyay, S., Rudolph, M. L., & Romanowicz, B. 2019,
GGG, 20, 4130
Yang, C. C., Johansen, A., & Carrera, D. 2017, A&A, 606, 80
Yang, C.-C., & Zhu, Z. 2021, MNRAS, 508, 5538
Yokochi, R., & Marty, B. 2004, E&PSL, 225, 77
Youdin, A. N. 2011, ApJ, 731, 99
Youdin, A. N., & Goodman, J. 2005, ApJ, 620, 459
Young, E. D., Kohl, I. E., Warren, P. H., et al. 2016, Sci, 351, 493
Zhang, J., Dauphas, N. M. D. A., Leya, I., & Fedkin, A. 2012, NatGe,
1429, 251
12
The Planetary Science Journal, 5:145 (12pp), 2024 June Nakajima et al.

Recommended for you

The Variable Detection of Atmospheric Escape around the Young, Hot Neptune AU...
The Variable Detection of Atmospheric Escape around the Young, Hot Neptune AU...The Variable Detection of Atmospheric Escape around the Young, Hot Neptune AU...
The Variable Detection of Atmospheric Escape around the Young, Hot Neptune AU...

This document summarizes observations from the Hubble Space Telescope of the young hot Neptune exoplanet AU Mic b, which orbits the nearby M dwarf star AU Mic. The observations aimed to detect atmospheric escape of neutral hydrogen through absorption in the stellar Lyman-alpha emission line. Two visits were obtained, one in 2020 and one in 2021, corresponding to transits of the planet. A stellar flare was observed and removed from the first visit data. In the second visit, absorption was detected in the blue wing of the Lyman-alpha line 2.5 hours before the white light transit, indicating the presence of high-velocity neutral hydrogen escaping the planet's atmosphere and traveling toward the observer. Estimates place the column density of this material

hubbleplanetatmosphere
Jack Oughton - Galaxy Formation Journal 02.doc
Jack Oughton - Galaxy Formation Journal 02.docJack Oughton - Galaxy Formation Journal 02.doc
Jack Oughton - Galaxy Formation Journal 02.doc

1) The document discusses theories of galaxy formation from the early universe following the Big Bang. 2) It describes the top-down and bottom-up theories of galaxy formation, where top-down suggests the first objects to form were large irregular structures that later broke apart, and bottom-up suggests smaller dense areas first combined together to form galaxies. 3) The author argues the bottom-up theory of smaller areas hierarchically clustering together is most credible given current evidence of hierarchical clustering in the universe, but knowledge in this area remains limited.

Large-scale Volcanism and the Heat Death of Terrestrial Worlds
Large-scale Volcanism and the Heat Death of Terrestrial WorldsLarge-scale Volcanism and the Heat Death of Terrestrial Worlds
Large-scale Volcanism and the Heat Death of Terrestrial Worlds

This document discusses the potential for large igneous provinces (LIPs) to cause the "heat death" of terrestrial planets through massive volcanic eruptions that overwhelm the climate system. It examines the timing of LIP events on Earth to estimate the likelihood of nearly simultaneous eruptions. Statistical analysis of Earth's LIP record finds that eruptions within 0.1-1 million years of each other are likely. Simultaneous LIPs could have driven Venus into a runaway greenhouse effect like its current state. The timing of LIP events on Earth provides insight into potential past LIP activity on Venus that may have ended its hypothesized earlier temperate climate.

venus

More Related Content

Similar to The Limited Role of the Streaming Instability during Moon and Exomoon Formation

Disks of Stars in the Galactic Center Triggered by Tidal Disruption Events
Disks of Stars in the Galactic Center Triggered by Tidal Disruption EventsDisks of Stars in the Galactic Center Triggered by Tidal Disruption Events
Disks of Stars in the Galactic Center Triggered by Tidal Disruption Events
Sérgio Sacani
 
The Exoplanet Radius Valley from Gas-driven Planet Migration and Breaking of ...
The Exoplanet Radius Valley from Gas-driven Planet Migration and Breaking of ...The Exoplanet Radius Valley from Gas-driven Planet Migration and Breaking of ...
The Exoplanet Radius Valley from Gas-driven Planet Migration and Breaking of ...
Sérgio Sacani
 
Highly siderophile elements_in_earth_mantle_as_a_clock_for_the_moon_forming_i...
Highly siderophile elements_in_earth_mantle_as_a_clock_for_the_moon_forming_i...Highly siderophile elements_in_earth_mantle_as_a_clock_for_the_moon_forming_i...
Highly siderophile elements_in_earth_mantle_as_a_clock_for_the_moon_forming_i...
Sérgio Sacani
 
Formation of low mass protostars and their circumstellar disks
Formation of low mass protostars and their circumstellar disksFormation of low mass protostars and their circumstellar disks
Formation of low mass protostars and their circumstellar disks
Sérgio Sacani
 
The debris of the ‘last major merger’ is dynamically young
The debris of the ‘last major merger’ is dynamically youngThe debris of the ‘last major merger’ is dynamically young
The debris of the ‘last major merger’ is dynamically young
Sérgio Sacani
 
on-the-origin-of-the-moon-and-the-continents.pdf
on-the-origin-of-the-moon-and-the-continents.pdfon-the-origin-of-the-moon-and-the-continents.pdf
on-the-origin-of-the-moon-and-the-continents.pdf
Medwin Publishers
 
A Tale of 3 Dwarf Planets: Ices and Organics on Sedna, Gonggong, and Quaoar f...
A Tale of 3 Dwarf Planets: Ices and Organics on Sedna, Gonggong, and Quaoar f...A Tale of 3 Dwarf Planets: Ices and Organics on Sedna, Gonggong, and Quaoar f...
A Tale of 3 Dwarf Planets: Ices and Organics on Sedna, Gonggong, and Quaoar f...
Sérgio Sacani
 
Younger moon
Younger moonYounger moon
Younger moon
Sérgio Sacani
 
Tidal star-planet interaction and its observed impact on stellar activity in ...
Tidal star-planet interaction and its observed impact on stellar activity in ...Tidal star-planet interaction and its observed impact on stellar activity in ...
Tidal star-planet interaction and its observed impact on stellar activity in ...
Sérgio Sacani
 
The Variable Detection of Atmospheric Escape around the Young, Hot Neptune AU...
The Variable Detection of Atmospheric Escape around the Young, Hot Neptune AU...The Variable Detection of Atmospheric Escape around the Young, Hot Neptune AU...
The Variable Detection of Atmospheric Escape around the Young, Hot Neptune AU...
Sérgio Sacani
 
Jack Oughton - Galaxy Formation Journal 02.doc
Jack Oughton - Galaxy Formation Journal 02.docJack Oughton - Galaxy Formation Journal 02.doc
Jack Oughton - Galaxy Formation Journal 02.doc
Jack Oughton
 
Large-scale Volcanism and the Heat Death of Terrestrial Worlds
Large-scale Volcanism and the Heat Death of Terrestrial WorldsLarge-scale Volcanism and the Heat Death of Terrestrial Worlds
Large-scale Volcanism and the Heat Death of Terrestrial Worlds
Sérgio Sacani
 
SO and SiS Emission Tracing an Embedded Planet and Compact 12CO and 13CO Coun...
SO and SiS Emission Tracing an Embedded Planet and Compact 12CO and 13CO Coun...SO and SiS Emission Tracing an Embedded Planet and Compact 12CO and 13CO Coun...
SO and SiS Emission Tracing an Embedded Planet and Compact 12CO and 13CO Coun...
Sérgio Sacani
 
Exomoons & Exorings with the Habitable Worlds Observatory I: On the Detection...
Exomoons & Exorings with the Habitable Worlds Observatory I: On the Detection...Exomoons & Exorings with the Habitable Worlds Observatory I: On the Detection...
Exomoons & Exorings with the Habitable Worlds Observatory I: On the Detection...
Sérgio Sacani
 
Hot Earth or Young Venus? A nearby transiting rocky planet mystery
Hot Earth or Young Venus? A nearby transiting rocky planet mysteryHot Earth or Young Venus? A nearby transiting rocky planet mystery
Hot Earth or Young Venus? A nearby transiting rocky planet mystery
Sérgio Sacani
 
On the theory_and_future_cosmic_planet_formation
On the theory_and_future_cosmic_planet_formationOn the theory_and_future_cosmic_planet_formation
On the theory_and_future_cosmic_planet_formation
Sérgio Sacani
 
N.36 turrini jupiter-and-the-early-collisional-evolution-of
N.36 turrini jupiter-and-the-early-collisional-evolution-ofN.36 turrini jupiter-and-the-early-collisional-evolution-of
N.36 turrini jupiter-and-the-early-collisional-evolution-of
IAPS
 
The Expansion of the X-Ray Nebula Around η Car
The Expansion of the X-Ray Nebula Around η CarThe Expansion of the X-Ray Nebula Around η Car
The Expansion of the X-Ray Nebula Around η Car
Sérgio Sacani
 
One tenth solar_abundances_along_the_body_of-the_stream
One tenth solar_abundances_along_the_body_of-the_streamOne tenth solar_abundances_along_the_body_of-the_stream
One tenth solar_abundances_along_the_body_of-the_stream
Sérgio Sacani
 
Lunar ejecta origin of near-Earth asteroid Kamo’oalewa is compatible with rar...
Lunar ejecta origin of near-Earth asteroid Kamo’oalewa is compatible with rar...Lunar ejecta origin of near-Earth asteroid Kamo’oalewa is compatible with rar...
Lunar ejecta origin of near-Earth asteroid Kamo’oalewa is compatible with rar...
Sérgio Sacani
 

Similar to The Limited Role of the Streaming Instability during Moon and Exomoon Formation (20)

Disks of Stars in the Galactic Center Triggered by Tidal Disruption Events
Disks of Stars in the Galactic Center Triggered by Tidal Disruption EventsDisks of Stars in the Galactic Center Triggered by Tidal Disruption Events
Disks of Stars in the Galactic Center Triggered by Tidal Disruption Events
 
The Exoplanet Radius Valley from Gas-driven Planet Migration and Breaking of ...
The Exoplanet Radius Valley from Gas-driven Planet Migration and Breaking of ...The Exoplanet Radius Valley from Gas-driven Planet Migration and Breaking of ...
The Exoplanet Radius Valley from Gas-driven Planet Migration and Breaking of ...
 
Highly siderophile elements_in_earth_mantle_as_a_clock_for_the_moon_forming_i...
Highly siderophile elements_in_earth_mantle_as_a_clock_for_the_moon_forming_i...Highly siderophile elements_in_earth_mantle_as_a_clock_for_the_moon_forming_i...
Highly siderophile elements_in_earth_mantle_as_a_clock_for_the_moon_forming_i...
 
Formation of low mass protostars and their circumstellar disks
Formation of low mass protostars and their circumstellar disksFormation of low mass protostars and their circumstellar disks
Formation of low mass protostars and their circumstellar disks
 
The debris of the ‘last major merger’ is dynamically young
The debris of the ‘last major merger’ is dynamically youngThe debris of the ‘last major merger’ is dynamically young
The debris of the ‘last major merger’ is dynamically young
 
on-the-origin-of-the-moon-and-the-continents.pdf
on-the-origin-of-the-moon-and-the-continents.pdfon-the-origin-of-the-moon-and-the-continents.pdf
on-the-origin-of-the-moon-and-the-continents.pdf
 
A Tale of 3 Dwarf Planets: Ices and Organics on Sedna, Gonggong, and Quaoar f...
A Tale of 3 Dwarf Planets: Ices and Organics on Sedna, Gonggong, and Quaoar f...A Tale of 3 Dwarf Planets: Ices and Organics on Sedna, Gonggong, and Quaoar f...
A Tale of 3 Dwarf Planets: Ices and Organics on Sedna, Gonggong, and Quaoar f...
 
Younger moon
Younger moonYounger moon
Younger moon
 
Tidal star-planet interaction and its observed impact on stellar activity in ...
Tidal star-planet interaction and its observed impact on stellar activity in ...Tidal star-planet interaction and its observed impact on stellar activity in ...
Tidal star-planet interaction and its observed impact on stellar activity in ...
 
The Variable Detection of Atmospheric Escape around the Young, Hot Neptune AU...
The Variable Detection of Atmospheric Escape around the Young, Hot Neptune AU...The Variable Detection of Atmospheric Escape around the Young, Hot Neptune AU...
The Variable Detection of Atmospheric Escape around the Young, Hot Neptune AU...
 
Jack Oughton - Galaxy Formation Journal 02.doc
Jack Oughton - Galaxy Formation Journal 02.docJack Oughton - Galaxy Formation Journal 02.doc
Jack Oughton - Galaxy Formation Journal 02.doc
 
Large-scale Volcanism and the Heat Death of Terrestrial Worlds
Large-scale Volcanism and the Heat Death of Terrestrial WorldsLarge-scale Volcanism and the Heat Death of Terrestrial Worlds
Large-scale Volcanism and the Heat Death of Terrestrial Worlds
 
SO and SiS Emission Tracing an Embedded Planet and Compact 12CO and 13CO Coun...
SO and SiS Emission Tracing an Embedded Planet and Compact 12CO and 13CO Coun...SO and SiS Emission Tracing an Embedded Planet and Compact 12CO and 13CO Coun...
SO and SiS Emission Tracing an Embedded Planet and Compact 12CO and 13CO Coun...
 
Exomoons & Exorings with the Habitable Worlds Observatory I: On the Detection...
Exomoons & Exorings with the Habitable Worlds Observatory I: On the Detection...Exomoons & Exorings with the Habitable Worlds Observatory I: On the Detection...
Exomoons & Exorings with the Habitable Worlds Observatory I: On the Detection...
 
Hot Earth or Young Venus? A nearby transiting rocky planet mystery
Hot Earth or Young Venus? A nearby transiting rocky planet mysteryHot Earth or Young Venus? A nearby transiting rocky planet mystery
Hot Earth or Young Venus? A nearby transiting rocky planet mystery
 
On the theory_and_future_cosmic_planet_formation
On the theory_and_future_cosmic_planet_formationOn the theory_and_future_cosmic_planet_formation
On the theory_and_future_cosmic_planet_formation
 
N.36 turrini jupiter-and-the-early-collisional-evolution-of
N.36 turrini jupiter-and-the-early-collisional-evolution-ofN.36 turrini jupiter-and-the-early-collisional-evolution-of
N.36 turrini jupiter-and-the-early-collisional-evolution-of
 
The Expansion of the X-Ray Nebula Around η Car
The Expansion of the X-Ray Nebula Around η CarThe Expansion of the X-Ray Nebula Around η Car
The Expansion of the X-Ray Nebula Around η Car
 
One tenth solar_abundances_along_the_body_of-the_stream
One tenth solar_abundances_along_the_body_of-the_streamOne tenth solar_abundances_along_the_body_of-the_stream
One tenth solar_abundances_along_the_body_of-the_stream
 
Lunar ejecta origin of near-Earth asteroid Kamo’oalewa is compatible with rar...
Lunar ejecta origin of near-Earth asteroid Kamo’oalewa is compatible with rar...Lunar ejecta origin of near-Earth asteroid Kamo’oalewa is compatible with rar...
Lunar ejecta origin of near-Earth asteroid Kamo’oalewa is compatible with rar...
 

More from Sérgio Sacani

A Strong He II λ1640 Emitter with an Extremely Blue UV Spectral Slope at z=8....
A Strong He II λ1640 Emitter with an Extremely Blue UV Spectral Slope at z=8....A Strong He II λ1640 Emitter with an Extremely Blue UV Spectral Slope at z=8....
A Strong He II λ1640 Emitter with an Extremely Blue UV Spectral Slope at z=8....
Sérgio Sacani
 
INEVITABLE ENDGAME OF COMET TSUCHINSHAN-ATLAS (C/2023 A3)
INEVITABLE ENDGAME OF COMET TSUCHINSHAN-ATLAS (C/2023 A3)INEVITABLE ENDGAME OF COMET TSUCHINSHAN-ATLAS (C/2023 A3)
INEVITABLE ENDGAME OF COMET TSUCHINSHAN-ATLAS (C/2023 A3)
Sérgio Sacani
 
The Dynamical Origins of the Dark Comets and a Proposed Evolutionary Track
The Dynamical Origins of the Dark Comets and a Proposed Evolutionary TrackThe Dynamical Origins of the Dark Comets and a Proposed Evolutionary Track
The Dynamical Origins of the Dark Comets and a Proposed Evolutionary Track
Sérgio Sacani
 
Possible Anthropogenic Contributions to the LAMP-observed Surficial Icy Regol...
Possible Anthropogenic Contributions to the LAMP-observed Surficial Icy Regol...Possible Anthropogenic Contributions to the LAMP-observed Surficial Icy Regol...
Possible Anthropogenic Contributions to the LAMP-observed Surficial Icy Regol...
Sérgio Sacani
 
Lunar Mobility Drivers and Needs - Artemis
Lunar Mobility Drivers and Needs - ArtemisLunar Mobility Drivers and Needs - Artemis
Lunar Mobility Drivers and Needs - Artemis
Sérgio Sacani
 
Transmission Spectroscopy of the Habitable Zone Exoplanet LHS 1140 b with JWS...
Transmission Spectroscopy of the Habitable Zone Exoplanet LHS 1140 b with JWS...Transmission Spectroscopy of the Habitable Zone Exoplanet LHS 1140 b with JWS...
Transmission Spectroscopy of the Habitable Zone Exoplanet LHS 1140 b with JWS...
Sérgio Sacani
 
Hydrogen sulfide and metal-enriched atmosphere for a Jupiter-mass exoplanet
Hydrogen sulfide and metal-enriched atmosphere for a Jupiter-mass exoplanetHydrogen sulfide and metal-enriched atmosphere for a Jupiter-mass exoplanet
Hydrogen sulfide and metal-enriched atmosphere for a Jupiter-mass exoplanet
Sérgio Sacani
 
The cryptoterrestrial hypothesis: A case for scientific openness to a conceal...
The cryptoterrestrial hypothesis: A case for scientific openness to a conceal...The cryptoterrestrial hypothesis: A case for scientific openness to a conceal...
The cryptoterrestrial hypothesis: A case for scientific openness to a conceal...
Sérgio Sacani
 
A slightly oblate dark matter halo revealed by a retrograde precessing Galact...
A slightly oblate dark matter halo revealed by a retrograde precessing Galact...A slightly oblate dark matter halo revealed by a retrograde precessing Galact...
A slightly oblate dark matter halo revealed by a retrograde precessing Galact...
Sérgio Sacani
 
A mature quasar at cosmic dawn revealed by JWST rest-frame infrared spectroscopy
A mature quasar at cosmic dawn revealed by JWST rest-frame infrared spectroscopyA mature quasar at cosmic dawn revealed by JWST rest-frame infrared spectroscopy
A mature quasar at cosmic dawn revealed by JWST rest-frame infrared spectroscopy
Sérgio Sacani
 
Search for Dark Matter Ionization on the Night Side of Jupiter with Cassini
Search for Dark Matter Ionization on the Night Side of Jupiter with CassiniSearch for Dark Matter Ionization on the Night Side of Jupiter with Cassini
Search for Dark Matter Ionization on the Night Side of Jupiter with Cassini
Sérgio Sacani
 
The X‐Pattern Merging of the Equatorial IonizationAnomaly Crests During Geoma...
The X‐Pattern Merging of the Equatorial IonizationAnomaly Crests During Geoma...The X‐Pattern Merging of the Equatorial IonizationAnomaly Crests During Geoma...
The X‐Pattern Merging of the Equatorial IonizationAnomaly Crests During Geoma...
Sérgio Sacani
 
The extremotolerant desert moss Syntrichia caninervis is a promising pioneer ...
The extremotolerant desert moss Syntrichia caninervis is a promising pioneer ...The extremotolerant desert moss Syntrichia caninervis is a promising pioneer ...
The extremotolerant desert moss Syntrichia caninervis is a promising pioneer ...
Sérgio Sacani
 
Measuring gravitational attraction with a lattice atom interferometer
Measuring gravitational attraction with a lattice atom interferometerMeasuring gravitational attraction with a lattice atom interferometer
Measuring gravitational attraction with a lattice atom interferometer
Sérgio Sacani
 
Discovery of Merging Twin Quasars at z=6.05
Discovery of Merging Twin Quasars at z=6.05Discovery of Merging Twin Quasars at z=6.05
Discovery of Merging Twin Quasars at z=6.05
Sérgio Sacani
 
Mapping the Growth of Supermassive Black Holes as a Function of Galaxy Stella...
Mapping the Growth of Supermassive Black Holes as a Function of Galaxy Stella...Mapping the Growth of Supermassive Black Holes as a Function of Galaxy Stella...
Mapping the Growth of Supermassive Black Holes as a Function of Galaxy Stella...
Sérgio Sacani
 
Compositions of iron-meteorite parent bodies constrainthe structure of the pr...
Compositions of iron-meteorite parent bodies constrainthe structure of the pr...Compositions of iron-meteorite parent bodies constrainthe structure of the pr...
Compositions of iron-meteorite parent bodies constrainthe structure of the pr...
Sérgio Sacani
 
Signatures of wave erosion in Titan’s coasts
Signatures of wave erosion in Titan’s coastsSignatures of wave erosion in Titan’s coasts
Signatures of wave erosion in Titan’s coasts
Sérgio Sacani
 
SDSS1335+0728: The awakening of a ∼ 106M⊙ black hole⋆
SDSS1335+0728: The awakening of a ∼ 106M⊙ black hole⋆SDSS1335+0728: The awakening of a ∼ 106M⊙ black hole⋆
SDSS1335+0728: The awakening of a ∼ 106M⊙ black hole⋆
Sérgio Sacani
 
Discovery of An Apparent Red, High-Velocity Type Ia Supernova at 𝐳 = 2.9 wi...
Discovery of An Apparent Red, High-Velocity Type Ia Supernova at  𝐳 = 2.9  wi...Discovery of An Apparent Red, High-Velocity Type Ia Supernova at  𝐳 = 2.9  wi...
Discovery of An Apparent Red, High-Velocity Type Ia Supernova at 𝐳 = 2.9 wi...
Sérgio Sacani
 

More from Sérgio Sacani (20)

A Strong He II λ1640 Emitter with an Extremely Blue UV Spectral Slope at z=8....
A Strong He II λ1640 Emitter with an Extremely Blue UV Spectral Slope at z=8....A Strong He II λ1640 Emitter with an Extremely Blue UV Spectral Slope at z=8....
A Strong He II λ1640 Emitter with an Extremely Blue UV Spectral Slope at z=8....
 
INEVITABLE ENDGAME OF COMET TSUCHINSHAN-ATLAS (C/2023 A3)
INEVITABLE ENDGAME OF COMET TSUCHINSHAN-ATLAS (C/2023 A3)INEVITABLE ENDGAME OF COMET TSUCHINSHAN-ATLAS (C/2023 A3)
INEVITABLE ENDGAME OF COMET TSUCHINSHAN-ATLAS (C/2023 A3)
 
The Dynamical Origins of the Dark Comets and a Proposed Evolutionary Track
The Dynamical Origins of the Dark Comets and a Proposed Evolutionary TrackThe Dynamical Origins of the Dark Comets and a Proposed Evolutionary Track
The Dynamical Origins of the Dark Comets and a Proposed Evolutionary Track
 
Possible Anthropogenic Contributions to the LAMP-observed Surficial Icy Regol...
Possible Anthropogenic Contributions to the LAMP-observed Surficial Icy Regol...Possible Anthropogenic Contributions to the LAMP-observed Surficial Icy Regol...
Possible Anthropogenic Contributions to the LAMP-observed Surficial Icy Regol...
 
Lunar Mobility Drivers and Needs - Artemis
Lunar Mobility Drivers and Needs - ArtemisLunar Mobility Drivers and Needs - Artemis
Lunar Mobility Drivers and Needs - Artemis
 
Transmission Spectroscopy of the Habitable Zone Exoplanet LHS 1140 b with JWS...
Transmission Spectroscopy of the Habitable Zone Exoplanet LHS 1140 b with JWS...Transmission Spectroscopy of the Habitable Zone Exoplanet LHS 1140 b with JWS...
Transmission Spectroscopy of the Habitable Zone Exoplanet LHS 1140 b with JWS...
 
Hydrogen sulfide and metal-enriched atmosphere for a Jupiter-mass exoplanet
Hydrogen sulfide and metal-enriched atmosphere for a Jupiter-mass exoplanetHydrogen sulfide and metal-enriched atmosphere for a Jupiter-mass exoplanet
Hydrogen sulfide and metal-enriched atmosphere for a Jupiter-mass exoplanet
 
The cryptoterrestrial hypothesis: A case for scientific openness to a conceal...
The cryptoterrestrial hypothesis: A case for scientific openness to a conceal...The cryptoterrestrial hypothesis: A case for scientific openness to a conceal...
The cryptoterrestrial hypothesis: A case for scientific openness to a conceal...
 
A slightly oblate dark matter halo revealed by a retrograde precessing Galact...
A slightly oblate dark matter halo revealed by a retrograde precessing Galact...A slightly oblate dark matter halo revealed by a retrograde precessing Galact...
A slightly oblate dark matter halo revealed by a retrograde precessing Galact...
 
A mature quasar at cosmic dawn revealed by JWST rest-frame infrared spectroscopy
A mature quasar at cosmic dawn revealed by JWST rest-frame infrared spectroscopyA mature quasar at cosmic dawn revealed by JWST rest-frame infrared spectroscopy
A mature quasar at cosmic dawn revealed by JWST rest-frame infrared spectroscopy
 
Search for Dark Matter Ionization on the Night Side of Jupiter with Cassini
Search for Dark Matter Ionization on the Night Side of Jupiter with CassiniSearch for Dark Matter Ionization on the Night Side of Jupiter with Cassini
Search for Dark Matter Ionization on the Night Side of Jupiter with Cassini
 
The X‐Pattern Merging of the Equatorial IonizationAnomaly Crests During Geoma...
The X‐Pattern Merging of the Equatorial IonizationAnomaly Crests During Geoma...The X‐Pattern Merging of the Equatorial IonizationAnomaly Crests During Geoma...
The X‐Pattern Merging of the Equatorial IonizationAnomaly Crests During Geoma...
 
The extremotolerant desert moss Syntrichia caninervis is a promising pioneer ...
The extremotolerant desert moss Syntrichia caninervis is a promising pioneer ...The extremotolerant desert moss Syntrichia caninervis is a promising pioneer ...
The extremotolerant desert moss Syntrichia caninervis is a promising pioneer ...
 
Measuring gravitational attraction with a lattice atom interferometer
Measuring gravitational attraction with a lattice atom interferometerMeasuring gravitational attraction with a lattice atom interferometer
Measuring gravitational attraction with a lattice atom interferometer
 
Discovery of Merging Twin Quasars at z=6.05
Discovery of Merging Twin Quasars at z=6.05Discovery of Merging Twin Quasars at z=6.05
Discovery of Merging Twin Quasars at z=6.05
 
Mapping the Growth of Supermassive Black Holes as a Function of Galaxy Stella...
Mapping the Growth of Supermassive Black Holes as a Function of Galaxy Stella...Mapping the Growth of Supermassive Black Holes as a Function of Galaxy Stella...
Mapping the Growth of Supermassive Black Holes as a Function of Galaxy Stella...
 
Compositions of iron-meteorite parent bodies constrainthe structure of the pr...
Compositions of iron-meteorite parent bodies constrainthe structure of the pr...Compositions of iron-meteorite parent bodies constrainthe structure of the pr...
Compositions of iron-meteorite parent bodies constrainthe structure of the pr...
 
Signatures of wave erosion in Titan’s coasts
Signatures of wave erosion in Titan’s coastsSignatures of wave erosion in Titan’s coasts
Signatures of wave erosion in Titan’s coasts
 
SDSS1335+0728: The awakening of a ∼ 106M⊙ black hole⋆
SDSS1335+0728: The awakening of a ∼ 106M⊙ black hole⋆SDSS1335+0728: The awakening of a ∼ 106M⊙ black hole⋆
SDSS1335+0728: The awakening of a ∼ 106M⊙ black hole⋆
 
Discovery of An Apparent Red, High-Velocity Type Ia Supernova at 𝐳 = 2.9 wi...
Discovery of An Apparent Red, High-Velocity Type Ia Supernova at  𝐳 = 2.9  wi...Discovery of An Apparent Red, High-Velocity Type Ia Supernova at  𝐳 = 2.9  wi...
Discovery of An Apparent Red, High-Velocity Type Ia Supernova at 𝐳 = 2.9 wi...
 

Recently uploaded

SUBJECT SPECIFIC ETHICAL ISSUES IN STUDY
SUBJECT SPECIFIC ETHICAL ISSUES IN STUDYSUBJECT SPECIFIC ETHICAL ISSUES IN STUDY
SUBJECT SPECIFIC ETHICAL ISSUES IN STUDY
Dr Kirpa Ram Jangra
 
Testing the Son of God Hypothesis (Jesus Christ)
Testing the Son of God Hypothesis (Jesus Christ)Testing the Son of God Hypothesis (Jesus Christ)
Testing the Son of God Hypothesis (Jesus Christ)
Robert Luk
 
Science grade 09 Lesson1-2 NLC-pptx.pptx
Science grade 09 Lesson1-2 NLC-pptx.pptxScience grade 09 Lesson1-2 NLC-pptx.pptx
Science grade 09 Lesson1-2 NLC-pptx.pptx
JoanaBanasen1
 
Properties of virus(Ultrastructure and types of virus)
Properties of virus(Ultrastructure and types of virus)Properties of virus(Ultrastructure and types of virus)
Properties of virus(Ultrastructure and types of virus)
saloniswain225
 
Keys of Identification for Indian Wood: A Seminar Report
Keys of Identification for Indian Wood: A Seminar ReportKeys of Identification for Indian Wood: A Seminar Report
Keys of Identification for Indian Wood: A Seminar Report
Gurjant Singh
 
Probing the northern Kaapvaal craton root with mantle-derived xenocrysts from...
Probing the northern Kaapvaal craton root with mantle-derived xenocrysts from...Probing the northern Kaapvaal craton root with mantle-derived xenocrysts from...
Probing the northern Kaapvaal craton root with mantle-derived xenocrysts from...
James AH Campbell
 
degree Certificate of Aston University
degree Certificate of Aston Universitydegree Certificate of Aston University
degree Certificate of Aston University
ebgyz
 
Anatomy, and reproduction of Gnetum.pptx
Anatomy, and reproduction of Gnetum.pptxAnatomy, and reproduction of Gnetum.pptx
Anatomy, and reproduction of Gnetum.pptx
karthiksaran8
 
Deploying DAPHNE Computational Intelligence on EuroHPC Vega for Benchmarking ...
Deploying DAPHNE Computational Intelligence on EuroHPC Vega for Benchmarking ...Deploying DAPHNE Computational Intelligence on EuroHPC Vega for Benchmarking ...
Deploying DAPHNE Computational Intelligence on EuroHPC Vega for Benchmarking ...
University of Maribor
 
SlideEgg-703870-Bird Migration PPT Presentation.pptx
SlideEgg-703870-Bird Migration PPT Presentation.pptxSlideEgg-703870-Bird Migration PPT Presentation.pptx
SlideEgg-703870-Bird Migration PPT Presentation.pptx
randaalmabrouk
 
Unveiling the Stability of Supermassive Black Hole Spin: Principal Component ...
Unveiling the Stability of Supermassive Black Hole Spin: Principal Component ...Unveiling the Stability of Supermassive Black Hole Spin: Principal Component ...
Unveiling the Stability of Supermassive Black Hole Spin: Principal Component ...
Ashkbiz Danehkar
 
Bragg Brentano Alignment for D4 with LynxEye Rev3.pptx
Bragg Brentano Alignment for D4 with LynxEye Rev3.pptxBragg Brentano Alignment for D4 with LynxEye Rev3.pptx
Bragg Brentano Alignment for D4 with LynxEye Rev3.pptx
Lisandro Cunci
 
TOPIC: INTRODUCTION TO FORENSIC SCIENCE.pptx
TOPIC: INTRODUCTION TO FORENSIC SCIENCE.pptxTOPIC: INTRODUCTION TO FORENSIC SCIENCE.pptx
TOPIC: INTRODUCTION TO FORENSIC SCIENCE.pptx
imansiipandeyy
 
The National Research Platform Enables a Growing Diversity of Users and Appl...
The National Research Platform Enables a Growing Diversity of Users and Appl...The National Research Platform Enables a Growing Diversity of Users and Appl...
The National Research Platform Enables a Growing Diversity of Users and Appl...
Larry Smarr
 
lipids_233455668899076544553879848657.pptx
lipids_233455668899076544553879848657.pptxlipids_233455668899076544553879848657.pptx
lipids_233455668899076544553879848657.pptx
muralinath2
 
Protein Digestion123334444556678890.pptx
Protein Digestion123334444556678890.pptxProtein Digestion123334444556678890.pptx
Protein Digestion123334444556678890.pptx
muralinath2
 
GIT hormones- II_12345677809876543235780963.pptx
GIT hormones- II_12345677809876543235780963.pptxGIT hormones- II_12345677809876543235780963.pptx
GIT hormones- II_12345677809876543235780963.pptx
muralinath2
 
Computer aided biopharmaceutical characterization
Computer aided biopharmaceutical characterizationComputer aided biopharmaceutical characterization
Computer aided biopharmaceutical characterization
souravpaul769171
 
Summer program introduction in Yunnan university
Summer program introduction in Yunnan universitySummer program introduction in Yunnan university
Summer program introduction in Yunnan university
Hayato Shimabukuro
 
Electrostatic force class 8 ncert. .pptx
Electrostatic force class 8 ncert. .pptxElectrostatic force class 8 ncert. .pptx
Electrostatic force class 8 ncert. .pptx
yokeswarikannan123
 

Recently uploaded (20)

SUBJECT SPECIFIC ETHICAL ISSUES IN STUDY
SUBJECT SPECIFIC ETHICAL ISSUES IN STUDYSUBJECT SPECIFIC ETHICAL ISSUES IN STUDY
SUBJECT SPECIFIC ETHICAL ISSUES IN STUDY
 
Testing the Son of God Hypothesis (Jesus Christ)
Testing the Son of God Hypothesis (Jesus Christ)Testing the Son of God Hypothesis (Jesus Christ)
Testing the Son of God Hypothesis (Jesus Christ)
 
Science grade 09 Lesson1-2 NLC-pptx.pptx
Science grade 09 Lesson1-2 NLC-pptx.pptxScience grade 09 Lesson1-2 NLC-pptx.pptx
Science grade 09 Lesson1-2 NLC-pptx.pptx
 
Properties of virus(Ultrastructure and types of virus)
Properties of virus(Ultrastructure and types of virus)Properties of virus(Ultrastructure and types of virus)
Properties of virus(Ultrastructure and types of virus)
 
Keys of Identification for Indian Wood: A Seminar Report
Keys of Identification for Indian Wood: A Seminar ReportKeys of Identification for Indian Wood: A Seminar Report
Keys of Identification for Indian Wood: A Seminar Report
 
Probing the northern Kaapvaal craton root with mantle-derived xenocrysts from...
Probing the northern Kaapvaal craton root with mantle-derived xenocrysts from...Probing the northern Kaapvaal craton root with mantle-derived xenocrysts from...
Probing the northern Kaapvaal craton root with mantle-derived xenocrysts from...
 
degree Certificate of Aston University
degree Certificate of Aston Universitydegree Certificate of Aston University
degree Certificate of Aston University
 
Anatomy, and reproduction of Gnetum.pptx
Anatomy, and reproduction of Gnetum.pptxAnatomy, and reproduction of Gnetum.pptx
Anatomy, and reproduction of Gnetum.pptx
 
Deploying DAPHNE Computational Intelligence on EuroHPC Vega for Benchmarking ...
Deploying DAPHNE Computational Intelligence on EuroHPC Vega for Benchmarking ...Deploying DAPHNE Computational Intelligence on EuroHPC Vega for Benchmarking ...
Deploying DAPHNE Computational Intelligence on EuroHPC Vega for Benchmarking ...
 
SlideEgg-703870-Bird Migration PPT Presentation.pptx
SlideEgg-703870-Bird Migration PPT Presentation.pptxSlideEgg-703870-Bird Migration PPT Presentation.pptx
SlideEgg-703870-Bird Migration PPT Presentation.pptx
 
Unveiling the Stability of Supermassive Black Hole Spin: Principal Component ...
Unveiling the Stability of Supermassive Black Hole Spin: Principal Component ...Unveiling the Stability of Supermassive Black Hole Spin: Principal Component ...
Unveiling the Stability of Supermassive Black Hole Spin: Principal Component ...
 
Bragg Brentano Alignment for D4 with LynxEye Rev3.pptx
Bragg Brentano Alignment for D4 with LynxEye Rev3.pptxBragg Brentano Alignment for D4 with LynxEye Rev3.pptx
Bragg Brentano Alignment for D4 with LynxEye Rev3.pptx
 
TOPIC: INTRODUCTION TO FORENSIC SCIENCE.pptx
TOPIC: INTRODUCTION TO FORENSIC SCIENCE.pptxTOPIC: INTRODUCTION TO FORENSIC SCIENCE.pptx
TOPIC: INTRODUCTION TO FORENSIC SCIENCE.pptx
 
The National Research Platform Enables a Growing Diversity of Users and Appl...
The National Research Platform Enables a Growing Diversity of Users and Appl...The National Research Platform Enables a Growing Diversity of Users and Appl...
The National Research Platform Enables a Growing Diversity of Users and Appl...
 
lipids_233455668899076544553879848657.pptx
lipids_233455668899076544553879848657.pptxlipids_233455668899076544553879848657.pptx
lipids_233455668899076544553879848657.pptx
 
Protein Digestion123334444556678890.pptx
Protein Digestion123334444556678890.pptxProtein Digestion123334444556678890.pptx
Protein Digestion123334444556678890.pptx
 
GIT hormones- II_12345677809876543235780963.pptx
GIT hormones- II_12345677809876543235780963.pptxGIT hormones- II_12345677809876543235780963.pptx
GIT hormones- II_12345677809876543235780963.pptx
 
Computer aided biopharmaceutical characterization
Computer aided biopharmaceutical characterizationComputer aided biopharmaceutical characterization
Computer aided biopharmaceutical characterization
 
Summer program introduction in Yunnan university
Summer program introduction in Yunnan universitySummer program introduction in Yunnan university
Summer program introduction in Yunnan university
 
Electrostatic force class 8 ncert. .pptx
Electrostatic force class 8 ncert. .pptxElectrostatic force class 8 ncert. .pptx
Electrostatic force class 8 ncert. .pptx
 

The Limited Role of the Streaming Instability during Moon and Exomoon Formation

  • 1. The Limited Role of the Streaming Instability during Moon and Exomoon Formation Miki Nakajima1,2 , Jeremy Atkins2 , Jacob B. Simon3 , and Alice C. Quillen2 1 Department of Earth and Environmental Sciences, University of Rochester, P.O. Box 270221, Rochester, NY 14627, USA 2 Department of Physics and Astronomy, University of Rochester, P.O. Box 270171, Rochester, NY 14627, USA 3 Department of Physics and Astronomy, Iowa State University, Ames, IA 50010, USA Received 2023 September 20; revised 2024 March 27; accepted 2024 April 17; published 2024 June 17 Abstract It is generally accepted that the Moon accreted from the disk formed by an impact between the proto-Earth and impactor, but its details are highly debated. Some models suggest that a Mars-sized impactor formed a silicate melt-rich (vapor-poor) disk around Earth, whereas other models suggest that a highly energetic impact produced a silicate vapor-rich disk. Such a vapor-rich disk, however, may not be suitable for the Moon formation, because moonlets, building blocks of the Moon, of 100 m–100 km in radius may experience strong gas drag and fall onto Earth on a short timescale, failing to grow further. This problem may be avoided if large moonlets (?100 km) form very quickly by streaming instability, which is a process to concentrate particles enough to cause gravitational collapse and rapid formation of planetesimals or moonlets. Here, we investigate the effect of the streaming instability in the Moon-forming disk for the first time and find that this instability can quickly form ∼100 km-sized moonlets. However, these moonlets are not large enough to avoid strong drag, and they still fall onto Earth quickly. This suggests that the vapor-rich disks may not form the large Moon, and therefore the models that produce vapor- poor disks are supported. This result is applicable to general impact-induced moon-forming disks, supporting the previous suggestion that small planets (<1.6 R⊕) are good candidates to host large moons because their impact- induced disks would likely be vapor-poor. We find a limited role of streaming instability in satellite formation in an impact-induced disk, whereas it plays a key role during planet formation. Unified Astronomy Thesaurus concepts: Earth-moon system (436) Supporting material: tar.gz file 1. Introduction 1.1. Review of the Moon-formation Hypotheses The origin of the Earth’s Moon is a long-standing open problem in planetary science. While it is accepted that the Moon formed from a partially vaporized disk generated by a collision between the proto-Earth and an impactor approxi- mately 4.5 billion yr ago (the “giant impact hypothesis”; Hartmann & Davis 1975; Cameron & Ward 1976), the details, such as the impactor radius and velocity, are actively debated. According to the canonical model, the proto-Earth was hit by a Mars-sized impactor (Canup & Asphaug 2001). This model can successfully explain several observations, such as the mass of the Moon, the angular momentum of the Earth–Moon system, and the small lunar core. Moreover, this model could potentially explain the observation that the Moon is depleted in volatiles, if they escaped during the impact or subsequent processes (see reviews by Halliday 2004; Canup et al. 2023). However, this model fails to explain the Earth and Moon’s nearly identical isotopic ratios (e.g., Si, Ti, W, and O; Armytage et al. 2012; Zhang & Dauphas 2012; Touboul et al. 2015; Thiemens et al. 2019, 2021; Kruijer et al. 2021). This is because the disk generated by an oblique Mars-sized impactor is primarily made of the impactor materials (Canup 2004; Kegerreis et al. 2022; Hull et al. 2024), which likely had different isotopic ratios from Earth, unless the inner solar system was well mixed and homogenized (Dauphas 2017) or equilibrated in the disk phase (Pahlevan & Stevenson 2007; Lock et al. 2018). In contrast, more energetic impact models may solve this problem by mixing the proto-Earth and impactor. These energetic models include an impact between two half-Earth-sized objects (Canup 2012), an impact between a rapidly rotating proto-Earth and a small impactor (Cúk & Stewart 2012), and general impact events that involve high energy and high angular momentum, which create a doughnut- shaped vapor disk connected to the Earth (a so-called synestia; Lock et al. 2018). In these scenarios, the disk and the proto- Earth compositions could have been mixed well, potentially solving the isotopic problem. Other proposed scenarios include (a) the Moon formation by multiple small impacts (Rufu et al. 2017), (b) hit-and-run collisions that mix the proto-Earth and the Moon more than the canonical model does (Asphaug et al. 2021), and (c) an impact between the proto-Earth covered by a surface magma ocean and an impactor, which leads to more contribution from Earth to the disk than the canonical scenario (Hosono et al. 2019). All of these models have potential explanations for the isotopic similarity, but each model faces challenges to explain other constraints. For example, the half-Earth and synestia models predict a much higher angular momentum of the Earth– Moon system than that of today (Canup 2012; Cúk & Stewart 2012). The excess of the angular momentum may or may not be removed easily (Cúk & Stewart 2012; Cúk et al. 2016, 2021; Rufu et al. 2017; Rufu & Canup 2020; Ward et al. 2020). Another constraint may come from the Earth’s interior; the deep portion of the Earth still retains solar neon (Yokochi & Marty 2004; Williams & Mukhopadhyay 2019) and helium isotopic ratios (Williams et al. 2019), which are distinct from the rest of the mantle. These noble gases may have been The Planetary Science Journal, 5:145 (12pp), 2024 June https://doi.org/10.3847/PSJ/ad4863 © 2024. The Author(s). Published by the American Astronomical Society. 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
  • 2. delivered to Earth’s mantle during the planet formation phase. This suggests that Earth’s mantle developed chemical hetero- geneity during Earth’s accretion, when the protoplanetary disk was still present, and that the Earth’s mantle has never been completely mixed, even by the Moon-forming impact. Previous work suggests that the preservation of the mantle heterogeneity can be explained by the canonical model but not by the energetic models, because these energetic impacts tend to mix the mantle (Nakajima & Stevenson 2015). It should be noted, however, the possibility that the neon and helium were delivered to the Earth's core and have been slowly leaking into the mantle over time has been discussed (Bouhifd et al. 2020), which would eliminate the need to preserve the mantle heterogeneity. However, it is not clear if delivery of these noble gases to the core was efficient. Moreover, due to recent analytical capability, small isotopic differences between Earth and the Moon have been observed (e.g., K, O, W, V, Cr, and others; Wiechert et al. 2001; Kruijer et al. 2015; Touboul et al. 2015; Wang & Jacobsen 2016; Young et al. 2016; Thiemens et al. 2019; Nielsen et al. 2021; Sossi et al. 2018). Some isotopic difference, such as the Moon’s enrichment in heavy K isotopes compared to those of Earth, can be explained by isotopic fractionation due to liquid– vapor phase separation during the Moon accretion process (Wang & Jacobsen 2016; Nie & Dauphas 2019; Charnoz et al. 2021). Observed small oxygen isotopic differences between Earth and the Moon may suggest that the Moon still records the impactor component (Cano et al. 2020). This suggestion may not be compatible with the energetic impact models because these impacts are so energetic that Earth and the Moon would be efficiently mixed. Whether the proposed models for the lunar origin can explain the observation that the Moon is depleted in volatiles is actively debated (Canup et al. 2015, 2023; Dauphas et al. 2015, 2022; Nakajima & Stevenson 2018; Sossi et al. 2018; Nie & Dauphas 2019; Charnoz et al. 2021; Halliday & Canup 2022; Canup et al. 2023). 1.2. Gas Drag Problem in a Vapor-rich Disk Another key constraint that has not been discussed until recently is the vapor mass fraction (VMF) of the disk, which sensitively depends on the impact model. Less energetic impacts, such as the canonical model and the multiple impact model, produce relatively small VMFs (VMF ∼ 0.2 for the canonical model, Nakajima & Stevenson 2014, and ∼0.1–0.5 for the multiple impact model, Rufu et al. 2017), while more energetic models, such as the half-Earth model and synestia model, produce nearly pure vapor disks (VMF ∼0.8–1, Nakajima & Stevenson 2014). The VMF of the Moon-forming disk can significantly impact the Moon accretion process; if the initial Moon-forming disk is vapor-poor (liquid-rich), moonlets can quickly form by gravitational instability (GI) from the liquid layer in the disk midplane outside the Roche radius (Thompson & Stevenson 1988; Salmon & Canup 2012). These moonlets eventually accrete to the Moon in tens to hundreds of years (Thompson & Stevenson 1988; Salmon & Canup 2012; Lock et al. 2018). In contrast, an initially vapor-rich disk needs to cool until liquid droplets emerge before moonlet accretion begins. Growing moonlets in the vapor-rich disk experience strong gas drag from the vapor (Nakajima et al. 2022). The gas drag effect is strongest when the gas and moonlet are coupled to an adequate degree and is sensitive to the moonlet radius. It is strongest with a moonlet radius Rp of a few kilometers (Nakajima et al. 2022). In contrast, much smaller moonlets are completely coupled with the gas, and much larger moonlets are completely decoupled from the gas, experiencing a weaker gas drag effect. As a result, ∼kilometer-sized moonlets lose their angular momentum and inspiral onto the Earth within 1 day (Nakajima et al. 2022), a much shorter timescale than that required for lunar formation. This same problem was a major challenge to planet formation in the protoplanetary disk (Adachi et al. 1976; Weidenschilling 1977). Here, we quickly review the gas drag problem in the protoplanetary disk. The radial velocity of a particle in a disk is (see Equation (A6) for the derivation; see also Armitage 2010; Takeuchi & Lin 2002) ( ) v v v 2 , 1 r r f 1 ,g K f f 1 t h t t = - + - - where τf = Ωtf is the dimensionless stopping time (Armitage 2010), Ω is the Keplerian angular velocity, tf is the friction time (see further descriptions below), vr,g is the gas velocity, and η is the pressure gradient parameter, described as ⎜ ⎟ ⎛ ⎝ ⎞ ⎠ ( ) c v p r 1 2 ln ln , 2 s K 2 g h = - ¶ ¶ where cs is the sound speed, vK is the Keplerian velocity, pg is the gas pressure, and r is the radial distance. vr is largest when τf = 1. The friction time is the time until the particle and gas reach the same velocity and is defined as tf = mvrel/FD, where m is the particle mass, vrel is the relative azimuthal velocity between the gas and particle, and FD is the drag force. In the protoplanetary disk, the gas drag for a particle is generally written as F R v C D 2 g,0 p 2 rel 2 D pr = - , where CD is the gas drag coefficient, ρg,0 is the initial gas density at the midplane, and Rp is the particle radius. The gas drag coefficients are roughly (Armitage 2010) ⎧ ⎨ ⎩ ( ) ( ) ( ) ( ) C 24 Re , Re 1 Stokesregime 24 Re , 1 Re 800 transitionregime 0.44, Re 800 Newtonregime , 3 D 1 0.6 = < < < > - - where Re is the Reynolds number. Assuming that vrel ∼ ηvK (see Equation (A2)), ν ∼ csλ, where ν is the kinematic viscosity, λ is the mean free path ( k T d p 2 B 2 g,0 l = p , where kB is the Boltzmann constant, T is the temperature, d is the molecular diameter, and pg,0 is the gas pressure at the midplane), Re = 4.26 v Rp rel ~ n , which is in the transition regime, in the protoplanetary disk at 1 au, assuming T = 280 K, η = 0.002, d = 289 pm, Rp = 0.52 m (see discussion below), p RT m g,0 g,0 m = r , and mm = 0.002 kg mol–1 , where mm is the mean molecular weight and R is the gas constant. Here we also assume the following vertical distribution of the gas density ρg, ⎜ ⎟ ⎛ ⎝ ⎞ ⎠ ( ) z H exp 2 , 4 g g,0 2 2 r r = - where z is the vertical coordinate and H is the gas scale height. This leads to H g,0 2 g r = p S . We assume Σg = 17,000 kg m−2 and use the relationships of c RT m s m = , H = cs/Ω, p r ln ln g = ¶ ¶ 2 The Planetary Science Journal, 5:145 (12pp), 2024 June Nakajima et al.
  • 3. ( ) 3 2 b - + , and 1 2 b = . The orbital angular frequency is GM r3 W = , where G is the gravitational constant, M is the stellar mass and r is the distance from the star (for the Moon- forming disk, M is the Earth mass and r is the distance from Earth). These parameters are taken from previous work (Carrera et al. 2015; other values of β have been discussed, such as 3 7 b = ; Chiang & Youdin 2010). The dimensionless stopping time becomes ( ) t C R v C R r 8 3 8 3 . 5 f f D p p g,0 rel D p p g,0 t r r h r r = W = W = τf = 1 when Rp = 0.52 m. The particle density ρp = 3000 kg m−3 is assumed. The residence time of the particle at 1 au is 1 au/vr = 75 yr, where vr is the radial fall velocity of the particle (see Equation (A7); for simplicity, the radial gas velocity vr,g = 0 is assumed). Thus, an approximately meter- sized particle falls toward the Sun at 1 au within 80 yr, a timescale much shorter than the planet formation timescale (several tens of Myr). This is the so-called “meter barrier” problem, which was a major issue to explain planetary growth in a classical planet formation model (Adachi et al. 1976; Weidenschilling 1977). In contrast, for the vapor-rich Moon-forming disk, τf = 1 (Re 1010 > ) is achieved when Rp = 1.8 km, assuming ρg,0 = 40 kg m–3 , ρp = 3000 kg m–3 , r = 3 R⊕, η = 0.04, T = 5200 K (see Section 2.3 for justifications for these parameters), and d = 300 pm. The residence time, r/vr, is approximately 1 day (1.21 days), which is much shorter than the Moon-formation timescale of several tens of yr to 100 yr. This formation timescale is ultimately determined by the radiative cooling timescale, but it is model-dependent. Here we provide a very simple estimate; the timescale for radiative cooling is ( ) 10 M L C T r T 4 p disk 2 ph 4 ~ p s + D yr (this is also consistent with numerical work; Lock et al. 2018), where Mdisk is the disk mass, L is the latent heat, Cp is the specific heat, ΔT is the temperature change over time, σ is the Stefan– Boltzmann constant, and Tph is the photosphere temperature. Here, Mdisk ∼ 0.015 M⊕, L = 1.2 × 107 J kg–1 (Melosh 2007), Cp = 103 J K–1 kg–1 , ΔT = 2000 K, and Tph = 2000 K (Thompson & Stevenson 1988) are assumed. However, the actual Moon- formation timescale can be longer than this for several reasons, including additional heating due to viscous spreading (Thompson & Stevenson 1988; Charnoz & Michaut 2015) and radial material transport efficiency (e.g., Salmon & Canup 2012). A short Moon- formation timescale has been proposed (e.g., Mullen & Gam- mie 2020; Kegerreis et al. 2022), but the typical Moon-formation timescale has been estimated to range in 10–102 yr. The vertical settling velocity of condensing particles is v C R z settle 8 3 D p p 2 g = r r W (Armitage 2010). Assuming z ∼ H, the settling time for a particle with a radius off 2 is 0.22 days. Determining the collision history of moonlets requires conducting orbital dynamics simulations, but here we provide a rough estimate. A rough estimate of the mass-doubling time of the largest moonlet in the Moon-forming disk is typically ∼1 day (Salmon & Canup 2012). If a 2 km-sized moonlet mass doubles in a day, the radius change is very small (2 km × 21/3 = 2.5 km), and the gas drag effect remains strong on such a moonlet; this change does not prevent the moonlet from falling into Earth. However, the actual collision time can vary depending on the local concentration of particles, which needs a detailed future study. This gas drag effect is a problem with forming the Moon from an initially vapor-rich disk. The Moon can still form after most of the vapor condenses, but by that time a significant portion of the disk mass could be lost (Ida et al. 2020; Nakajima et al. 2022), which would fail to form a large moon from an initially vapor-rich disk (here we use a ”large” moon when its mass is ∼1 wt% or larger of the host planet). If this is the case, an initially vapor-rich disk may not be capable of forming a large Moon (Nakajima et al. 2022). In contrast, the gas drag effect is weak for vapor-poor disks, which are generated by less energetic models, such as the canonical and multiple impact models. 1.3. Streaming Instability in the General Impact-induced Moon-forming Disk A potential solution to this vapor drag problem is forming a large moonlet very quickly (much larger than 2 km), so that the moonlet would not experience strong gas drag. This is the accepted solution for the gas drag problem in the protoplane- tary disk. The proposed mechanism is the streaming instability (Youdin & Goodman 2005; Johansen et al. 2007), where particles spontaneously concentrate in the disk, gravitationally collapsing and forming a large clump (∼100 km in size; Johansen et al. 2015). If this mechanism works for the Moon- forming disk, an initially vapor-rich disk may be able to form the Moon despite the gas drag issue. If this mechanism turns out not to work for the Moon-forming disk, it is an interesting finding as well, given that a Moon-forming disk is often treated as a miniature analog of the protoplanetary disk. Understanding what makes these two disks differ would deepen our under- standing of planet and satellite formation processes. Moreover, knowing whether the streaming instability can affect moon formation processes informs our understanding of moon formation in the solar and extrasolar systems. Moon formation in an impact-induced disk is common in the solar system (e.g., the moons of Mars (Craddock 2011), the moons of Uranus (Slattery et al. 1992), and Pluto-Charon (Canup 2005)). While there are no confirmed exomoons (moons around exoplanets) to date (e.g., Cassese & Kipping 2022), impact-induced exomoons should be common because impacts are a common process during planet formation (Nakajima et al. 2022) and because impacts in extrasolar systems may have already been observed (e.g., Meng et al. 2014; Bonomo et al. 2019; Thompson et al. 2019; Kenworthy et al. 2023). If streaming instability operates in these disks, it can affect what types of planets can host exomoons, which can be compared with future exomoon observations. It should also be noted that other instability, such as secular GI (e.g., Youdin 2011; Takahashi & Inutsuka 2014; Tominaga et al. 2018) and two-component viscous GI (TVGI; Tominaga et al. 2019), have been discussed as a mechanism for planetesimal and dust ring formation. The secular GI occurs when dust–gas interaction reduces the rotational support of the rotating disk, which leads to dust concentration. The TVGI is an instability caused by dust–gas friction and turbulent gas viscosity. These instabilities can lead to clump formation even if the disk is self-gravitationally stable. The implications of these instabilities are beyond the scope of this paper. 3 The Planetary Science Journal, 5:145 (12pp), 2024 June Nakajima et al.
  • 4. 1.4. Motivation of This Work The goal of this work is to investigate whether the streaming instability can form moonlets that are large enough to avoid strong gas drag from the vapor-rich disk. The result could constrain the Moon-formation model as well as general impact- induced models in the solar and extrasolar systems. We conduct hydrodynamic simulations using the code Athena (Stone et al. 2008; Bai & Stone 2010a). We first conduct 2D simulations to identify the section of parameter space that leads to streaming instability. Subsequently, we conduct 3D simula- tions with self-gravity to identify the size distribution of the moonlets. Finally, we identify the lifetime of the moonlets formed by streaming instability to investigate whether it is possible to form a large moon from an initially vapor-rich disk. For the general impact-induced disks, we consider ”rocky” and ”icy” disks, where these disks form by collisions between rocky planets (with silicate mantles and iron cores) and between icy planets (with water-ice mantles and silicate cores), respectively. 2. Method 2.1. Athena We use the Athena hydrodynamics code, which solves the equations of hydrodynamics using a second-order accurate Godunov flux-conservative approach (Stone et al. 2008). We use the configuration of Athena that couples the dimensionally unsplit corner transport upwind method (Colella 1990) to the third-order in-space piecewise parabolic method by Colella & Woodward (1984) and calculates the numerical fluxes using the HLLC Riemann solver (Toro 1999). We also integrate the equations of motion for particles following Bai & Stone (2010a) and include particle self-gravity for 3D simulations following the particle-mesh approach described in previous work (Simon et al. 2016). Orbital advection is taken into account following previous work (Bai & Stone 2010a, 2010b). Our setup is the local shearing box approximation in which a small patch of the disk is corotating with the disk at the Keplerian velocity (Stone & Gardiner 2010). The local Cartesian frame is defined as (x, z) for 2D and (x, y, z) for 3D simulations, with x as the radial coordinate from the planet, z parallel to the planetary spin axis, and y in the direction of the orbital rotation. In the Athena code, we solve the following equations for our simulations (Bai & Stone 2010a; Simon et al. 2016; Li et al. 2019): · ( ) ( ) t u 0, 6 g g r r ¶ ¶ +  = · ( ) ( ) t p t u uu I x z u v u 3 2 , 7 g g g g 2 g 2 g p f r r r r r r W ¶ ¶ +  + = W - W - ´ + - ˆ ˆ ˆ ( ) d dt v x z t v x x z v v u a 2 3 2 . 8 i i i i i K 2 2 f g h W = - W + W -W - ´ - - + The first two equations specify mass and momentum conservation for the gas, respectively, while the third equation represents the motion of a particle i coupled with the gas. Here, u is the velocity of the gas and I is the identity matrix. v is the mass-weighted averaged particle velocity in the fluid element, assuming that particles can be treated as fluid (Bai & Stone 2010a). The terms of the right-hand side of Equation (7) are radial tidal forces (gravity and centrifugal force), vertical gravity, and the Coriolis force and the feedback from the particle to the gas. In Equation (8), the first term on the right-hand side describes a constant radial force due to gas drag. vi is the particle velocity, x̂ and ẑ represent the unit vectors in the x- and z-directions, and xi and zi are the values of x and z for the particle i. u is the gas velocity interpolated from the grid cell centers to the location of the particle. The second, third, and fourth terms are radial and vertical tidal forces and the Coriolis force. ag is the acceleration due to self-gravity, which is considered only in 3D simulations. ag = −∇Φp, where Φp is the gravitational potential and satisfies the Poisson’s equation, ∇2 Φp = 4πGρp. To reduce computational time, the particles are organized into “superparticles,” each representing a cluster of individual particles of the same size. In the code units, we normalize Ω = cs = H = ρg,0 = 1. The gas and particle initial distributions are described in Equation (4) and as ⎛ ⎝ ⎜ ⎞ ⎠ ⎟ ( ) H z H 2 exp 2 , 9 p p p 2 p 2 r p = S - respectively, where Hp is the scale height of the particles and is set to 0.02H (Bai & Stone 2010a) and Σp is the particle surface density. The system uses an isothermal equation of state P c g s 2 r = , and the particles are distributed uniformly in the x- and y-directions and normally in the z-direction (Equation (9)). The computational domains are 0.2H × 0.2H in 2D simulations and 0.2H × 0.2H × 0.2H in 3D simulations, with all-periodic boundary conditions. The resolution for 2D is 512 × 512, and each grid cell has one particle (512 × 512 = 262,144 particles). The resolution for 3D is 10 × 1283 (= 21 million particles in total). Previous work shows that the resolution of 1283 can produce the large clump size distribution well compared to those with 2563 and 5123 (Simon et al. 2016); therefore, this resolution of our 3D simulations is sufficient to resolve large clumps, which are the focus of this work. The initial particle size is assumed to be constant. Variable initial particle sizes could affect the growth speed (Krapp et al. 2019) and the concentrations of particles depending on their sizes (Yang & Zhu 2021), but it is not known to affect the largest clump size formed by streaming instability. 2.2. Clump Detection in 2D and 3D After we conduct 2D simulations, we identify filaments forming in the disk in order to constrain the parameter space favorable for the streaming instability. We use the Kolmo- gorov–Smirnov (KS) method, which is based on previous work (Carrera et al. 2015). For each time step in a given time window (25/Ω in our case), we compute the particle surface density ( ) ( ) x x z dz , p p 0.1 0.1 ò r S = - and average it over the time window to give p t áS ñ . We then sort the values in p t áS ñ from highest to lowest and compute the cumulative distribution of this sorted data set. Finally, we use the KS test to output a 4 The Planetary Science Journal, 5:145 (12pp), 2024 June Nakajima et al.
  • 5. p-value that measures the likelihood that the underlying cumulative distribution was linear: ( ) ( ) ( ) ( ) Q z e p Q D n 2 1 , 10 j j z j 1 1 2 2 2 å = - = = ¥ - - where D is the maximum distance between the data and linear cumulative distributions and n is the number of data points. The higher the p is, the more homogeneous the system is, and therefore filament formation is unlikely, whereas a small p indicates filament formation is more likely. Here we assume that filament formation is very likely when p < 0.10, the same as in previous work (Carrera et al. 2015). For self-gravitating clump detection in our 3D simulations, we use PLanetesimal ANalyzer (PLAN; Li et al. 2019). This is a tool to identify self-gravitating clumps specifically made for Athena output. The density of each particle is assessed based on the nearest particles. Particles with densities higher than a threshold are associated with neighboring dense particles until a density peak is achieved. In contrast, particle groups with a saddle point less than a threshold remain separated. Detailed descriptions are found in previous work (Li et al. 2019). 2.3. Model Parameters The main parameters for the Athena simulations are (1) the dimensionless stopping time τf (the Newton regime; see Section 1.2 and Nakajima et al. 2022), (2) the normalized pressure parameter Δ = ηvK/cs, (3) the ratio of the particle surface density to the gas surface density Z = Σp/Σg (VMF = Z 1 1 + ), and (4) the normalized gravity G̃ G 4 g,0 2 p r º W for 3D simulations, where G is the gravitational constant. A small value of τf indicates a small particle radius Rp, where the particle is well coupled with the gas, whereas a large value of τf corresponds to a large value of Rp, which is more decoupled from the gas. A large value of Δ corresponds to a large pressure gradient and quicker radial infall. G̃ represents the strength of the self-gravity. The parameter space we are exploring for our 2D simulations is τf = 10−3 , 10−2 , 10−1 , 100 , 101 ; Δ = 0.1, 0.2, 0.3, 0.4, 0.5; and Z = 0.05 and 0.1. For 3D simulations, we use Z = 0.1 and 0.3 and G̃ 0.1788 = and 0.5898, where the former G̃ value corresponds to a slightly cooler temperature (4700 K), while the latter corresponds to a hotter temperature (5200 K). Here, we justify the choice of these parameters. This range of τf corresponds to the particle radii between 2 m and 20 km (see Equation (5)). The global disk structures have been calculated based on hydrodynamic simula- tions in previous work. The overall disk mass of the Moon- forming disk is typically a few percent of Earth, depending on the impact model (MD/ML = 1.35–2.80, where MD is the disk mass and ML is the lunar mass; Nakajima & Stevenson 2014). The midplane disk temperature ranges from 3000 K to 7000 K, and the radial range of the disk is ∼1–8 R⊕ (see Figure 5 in Nakajima & Stevenson 2014). The disk temperature is ∼4000–5500 K at r = 3 R⊕. For the general rocky and icy impact-induced disks, the disk temperature can vary, typically in the range of thousands of K for vapor-rich disks (Nakajima et al. 2022). The pressure gradient can vary, but the typical value of η is ∼0.02–0.06 based on impact simulations (Nakajima et al. 2022). In the Moon-forming disk, the value of Z increases as the disk cools (on a timescale of 10–102 yr). In other words, Z can initially be zero in an energetic Moon-forming impact model (e.g., Lock et al. 2018) and eventually becomes infinity as the disk materials condense and the gas disappears. Thus, picking a value of Z means that we are seeing physics at a specific time. Since we are primarily interested in high-vapor disks (i.e., an early phase of the disk), we explore the mass ratio range Z ä [0.05, 0.1] for our 2D simulations and Z ä [0.1, 0.3] for our 3D simulations. We use the large value of Z = 0.3 as a sensitivity test. Higher values of Z can be achieved as the disk cools, but we focus on small values of Z („0.3) for several reasons. First, at a larger value of Z (>0.3), the conventional GI in the liquid part of the disk can form moonlets. At Z 0.76 Z 1 1 = + , and the total (Σp + Σg) surface density at r ∼ 3 R⊕ is ∼108 kg m−2 in energetic models, which means that Σp = 0.76 × 108 kg m−2 = 7.6 × 107 kg m−2 . Previous work suggests that when the liquid (melt) layer’s thickness reaches 5–10 km (or the equivalent surface density of a few 107 kg m−2 ), GI can happen in the melt layer (Machida & Abe 2004). Whether streaming instability occurs at the same time or whether streaming instability affects the GI in the Moon-forming disk are unknown and have not been explored in previous studies. Under these circumstances, the important of streaming instability becomes unclear. Additionally, these streaming instability simulations have been conducted at low-Z values („0.1) in previous work to reproduce conditions in the protoplanetary disks (e.g., Abod et al. 2019; Li & Youdin 2021), and simulations with high-Z (>0.3) values have not been fully tested. For these reasons, we focus on relatively small values of Z in this study. For the set of Athena simulations, we focus on reproducing two sets of the Moon-forming disk thermal profile; assuming M = M⊕ and r = 3 R⊕, where M⊕ is the Earth mass and R⊕ is the Earth radius, the gas pressure at the midplane around 3 R⊕ is ≈12 MPa, T = 4700 K, and mm = 30 g mol−1 . This makes the density ρg,0 = 12.13 kg m−3 for an ideal gas, cs = 1140 m s−1 , vK = 4562 m s−1 , Ω = 2.38 × 10−4 s−1 , the scale height H = cs/Ω = 4.78 × 106 m, and G̃ 0.1788 = . For the higher- temperature scenario, T = 5200 K, ρg,0 = 40.01 kg m−3 , cs = 1200 m s−1 , H = 5.03 × 106 , and G̃ 0.5898 = . These temperatures are motivated by previous hydrodynamic calcula- tions of the Moon-forming disk formation (Nakajima & Stevenson 2014), and the other parameters are calculated based on an equation of state of dunite assuming that the vapor and liquid phases are in equilibrium (MANEOS; Thompson & Lauson 1972; Melosh 2007). Since we are assuming that the disk is in liquid–vapor equilibrium, a higher disk temperature leads to a higher vapor density at the midplane. For example, the gas density at liquid–vapor equilibrium is 150 kg m−3 at 6000 K and 1.8 kg m−3 at 4000 K for dunite, according to MANEOS (Thompson & Lauson 1972; Melosh 2007). Assum- ing η ∼ 0.02–0.06 for the Moon-forming disk, Δ ∼ 0.1–0.2. However, higher values are possible (Nakajima et al. 2022); therefore, we explore the range of Δ ∼ 0.1–0.5. The parameter values of Δ and G̃ are significantly different from values in the protoplanetary disk, where the typical values used are Δ ∼ 10−3 –10−2 and G̃ 0.05 ~ in the protoplanetary disk (Carrera et al. 2015; Simon et al. 2016). We use a fixed value of Z in hydrodynamic calculations because the condensation timescale (∼years) is longer than the simulation timescale (we run 2D simulations for ∼100 orbits, which correspond to 123 hr. A steady state is reached by this time). 5 The Planetary Science Journal, 5:145 (12pp), 2024 June Nakajima et al.
  • 6. We also assume that the disk does not evolve on this short timescale. 3. Results 3.1. 2D Athena Simulations Figure 1(A) shows four examples of spacetime plots of our 2D simulations, where (1) (τf, Δ) = 10−3 , 0.3, (2) 10−2 , 0.1, (3) 10−1 , 0.2, and (4) 10−1 , 0.2. Z = 0.1 for the four cases. These simulations represent four characteristic regimes. The horizontal axis is x normalized by the scale height H, and the vertical axis is the number of orbits. The color shows particle concentration. In case 1, τf is small, which means that gas and particles are well coupled, and this does not lead to filament formation. In case 2, after ∼20 orbits, filaments start forming, and these filaments are stable during the rest of the simulations, which indicate that streaming instability occurs with this chosen parameter. A radially shearing periodic boundary condition is used in the x-direction; therefore, the filament that appears at x/H = −0.1 at ∼100 orbits reappears at x/H = 0.1. Two distinct filaments form in this simulation. In case 3, filaments form, but they are not stable because their radial movements are high due to the high Δ value. Thus, this is not an ideal parameter space for streaming instability. Similar behaviors have been seen in simulations for the protoplanetary disk with other parameter combinations (e.g., Δ = 0.05, τf = 3, Z = 0.005; Carrera et al. 2015). In case 4, the filaments are not as clearly defined as those in case 2, but a coherent filament structure is found after several orbits. Thus, this is also considered as a streaming instability regime. Figure 1(B) shows the summary of our 2D simulations for Z = 0.1 (top) and Z = 0.05 (bottom). We use the KS test to identify the extent of particle concentration (see Section 2.2). Concentration is measured by the p-value, and when the value of p is small (p < 0.1), streaming instability is likely (Carrera et al. 2015). The color bar indicates the p-value. For Z = 0.1, p < 0.1 is achieved when Δ = 0.1 and 10−3 „ τf „ 100 . For large Δ values, the radial motions are large, and stable filaments do not form, as discussed above. For Z = 0.05, this trend remains the same, but the streaming instability regime is smaller (10−2 „ τf „ 100 ) than that for Z = 0.1. This is because the larger Z leads to the presence of more particles in the disk and therefore to more filament formation (Carrera et al. 2015). Thus, our 2D simulations show that the most filaments form at relatively small Δ (Δ = 0.1) and with both Z = 0.1 and Z = 0.05, but the higher Z has more favorable conditions. This general trend is consistent with previous work in the protoplanetary disk (e.g., Carrera et al. 2015), which investigates smaller Δ values for the protoplanetary disk (Δ „ 0.05). 3.2. 3D Athena Simulations Now that we have identified the parameter space for the streaming instability (Δ = 0.1, Z = 0.1), we perform 3D simulations with self-gravity to estimate the size distributions of streaming-instability-induced moonlets. Figure 2 shows Figure 1. (A) Spacetime diagram for the four cases. The input parameters are (1) (τf, Δ) = 10−3 , 0.3, (2) 10−2 , 0.1, (3) 10−1 , 0.2, and (4) 1, 0.1, all at Z = 0.1. The horizontal axis is x normalized by the scale height H. The vertical axis indicates the number of orbits. The color shows p p S áS ñ, where Σp is the particle surface density and p áS ñ is the average along the x-axis. Filament formation occurs in cases 2 and 4, while no filament formation occurs in cases 1 and 3. (B) Summary of results for Z = 0.1 and Z = 0.05. The colors show the p-value, and the clumping regime (p < 0.1) is shown in the sky blue shade. Parameters for cases 1–4 are indicated. This shows that clumping occurs only at small Δ values (Δ = 0.1). 6 The Planetary Science Journal, 5:145 (12pp), 2024 June Nakajima et al.
  • 7. snapshots of one of our 3D simulations ( G̃ 1, f t = = Z 0.5898, 0.1 = , and Δ = 0.1) at four different times (t = 0.32, 2.87, 3.39, and 3.55), where t is the number of orbits. The self-gravity is not initially included and is turned on at t = 3.18. This is a general practice to avoid artificial clumping before streaming instability takes place (Simon et al. 2016). The color shows the particle density. The circles indicate the Hill spheres of self-bound clumps, which are identified using PLAN (Li et al. 2019). At t = 0.32, no clump is identified, but by t = 2.87, the streaming instability fully develops and reaches its steady state. After self-gravity is turned on at t = 3.18, self-gravitating clumps form right away (t = 3.39). This general behavior is similar to previous work on the streaming instability in the protoplanetary disk (Abod et al. 2019), but the time it takes to form clumps by streaming instability is shorter, probably because of the large Z value (Simon et al. 2022) compared to those of the protoplane- tary disk. Figure 3(A) shows the cumulative distribution of moonlet mass Mp, normalized by the characteristic self-gravitating mass MG, defined as (Abod et al. 2019) ⎛ ⎝ ⎞ ⎠ ˜ ( ) ( ) M G Z G H 2 4 2 2 , 11 G G p p 2 5 2 3 4 9 2 3 2 g,0 3 p l p p r = S = S W = where λG is an instability wavelength, which originates from the Toomre dispersion relation, equating the tidal and gravitational forces (Abod et al. 2019). The parameters are τf = 0.1, 1, G̃ 0.1788, 0.5898 = , and Z = 0.1, 0.3. For the G̃ 0.1787 = cases, MG = 5.19 × 1018 kg, and for the G̃ 0.5898 = cases, MG = 2.17 × 1020 kg. For all 3D simula- tions, Δ = 0.1 is assumed. The solid lines indicate the moonlet mass distribution shortly after the onset of the self-gravity (t = 3.66 for G̃ 0.1788, 0.1 f t = = , and t = 3.34 for all the other cases). The dashed–dotted lines indicate the same parameter after an additional time (t + Δt, where Δt = 1/ Ω = 0.16 orbits, except the Z = 0.3 case, where t 1 2 D = W ). The Figure 2. Snapshots of a 3D simulation (τf = 1, Δ = 0.1, Z = 0.1, G̃ 0.5898 = ). The horizontal and vertical axes are x and y, normalized by the scale height H. The color indicates the particle density, normalized by the average density. t indicates the number of orbits. The circles represent the Hill radius of each moonlet formed by streaming instability. At first (t = 0.32), no concentration of particles is observed, but streaming instability clearly develops by t = 2.87. After self-gravity is turned on at t = 3.18, moonlets form by GI (t = 3.39, 3.55). The self-gravitating clumps are identified using PLAN (see the main text). 7 The Planetary Science Journal, 5:145 (12pp), 2024 June Nakajima et al.
  • 8. reason for the shorter Δt for the higher-Z case is that the streaming instability develops faster for higher-Z values (Simon et al. 2022). The maximum Mp/MG = 0.254 (G̃ 0.1788, = 1 f t = at t = 3.50). This is broadly consistent with previous studies that suggest that the maximum clump masses formed by the streaming instability are characterized by MG in the range of 10−1 −101 MG (Abod et al. 2019; Li et al. 2019). Our result lies on the lower end of this mass range. This indicates that the moonlet mass distribution based on our numerical simulations is consistent with the analytical mass model generated for the protoplanetary disk. Thus, this also indicates that the processes of streaming instability are similar between the Moon-forming disk and protoplanetary disk despite the different input parameters. We also find that higher Z does not necessarily lead to higher Mp. Here we consider the best-case scenario of forming a large moonlet. Assuming that the density of the clump is 3000 kg m−3 , MG = 2.17 × 1020 kg is equivalent to 258 km in radius and 0.254 MG = 163 km. Based on Equation (1), the residence times of 258 km and 163 km moonlets are 92 and 58 days, respectively, assuming vr,g = 0. These timescales are much shorter than the Moon-formation timescale (10 s–100 yr, Thompson & Steven- son 1988). Thus, even though the streaming instability can occur in an initially vapor-rich Moon-forming disks, it does not help increase the residence time of moonlets. 3.3. Exomoon Formation by Streaming Instability The streaming instability likely occurs in impact-induced moon-forming disks in extrasolar systems. The pressure gradient parameter η (∼0.02–0.06) is similar regardless of the composition of the disk (Nakajima et al. 2022). Figure 3(B) shows the disk residence time for a clump with mass MG in a moon-forming disk as a function of the planetary mass (Mplanet = 1–6 M⊕). “Rocky planet” corresponds to disks formed by a collision between terrestrial planets, while “icy planet” indicates disks formed by a collision between icy planets whose mantles are made of water ice (70 wt%) and cores are made of iron (30 wt%). This is produced by calculating r/vr (see Equation (1)), where for rocky planets, we use ρg,0 = 40.01 kg m−3 , T = 5200 K, ρp = 3000 kg m−3 , and r = 3 Rplanet, where Rplanet is the planetary radius, and for icy planets we use ρg,0 = 10.0 kg m−3 , T = 2000 K, ρp = 1000 kg m−3 , and r = 3 Rplanet. These are the temperatures when the impact-induced disks reach complete vaporization based on impact simulations (Nakajima et al. 2022). A higher temperature is possible, but the disk would need to cool down to this temperature to form particles (dust); therefore, this is the most relevant temperature to assess the effect of streaming instability. The thermal state of the disk formed by icy planetary collisions is estimated based on an equation of state of water (Senft & Stewart 2008). Additionally, ( ) R R M M planet planet 1 3.7 = Å Å for rocky planets and ( ) R R M M 1.2 planet planet 1 3 = Å Å for icy planets are assumed (Kipping et al. 2013; Mordasini et al. 2012). At Mplanet = 1–6 M⊕, these parameters make G̃ 0.42 0.67 = - for rocky planets and G̃ 0.25 = for icy planets, which are similar to the ranges covered in our hydrodynamic simulations (Section 2.3). In both icy and rocky planet cases, the disk residence time is a few tens of days to several months, which is short compared to the satellite formation timescale (10 s–100 s yr, Nakajima et al. 2022). This suggests that the streaming instability likely plays a limited role in impact-induced moon-forming disks. 4. Discussion 4.1. Streaming Instability in the Moon-forming Disk Figure 4 shows a schematic view of our hypothesis. An energetic impact would generate a vapor-rich disk (the disk is made of silicate vapor for the Moon-forming disk and rocky moon-forming disks and water vapor for icy moon-forming disks). Over time, the disk cools by radiation and small droplets (<cm) emerge. These small droplets would grow by accretion and streaming instability. However, once these moonlets reach 100 m–100 km in size, gas drag from the vapor is so strong that they fall onto the planet on a short timescale (days–weeks). This continues to occur until the VMF of the disk decreases so that the gas drag effect is no longer strong (we assume the silicate vapor is in equilibrium with the silicate liquid. As the disk temperature decreases, the gas density decreases). Once this condition is reached, a liquid layer emerges, and moonlets can stay in the disk. However, by this time, a significant disk mass could have been lost (>80 wt%; Ida et al. 2020; Nakajima et al. 2022). For this reason, only a small moon (or moons) Figure 3. (A) Cumulative mass distribution of moonlets formed by streaming instability for Δ = 0.1, Z = 0.1, 0.3. The purple, dark blue, light blue, dark yellow, and light yellow lines indicate parameter values of ( ˜ ) G Z , , f t = (0.5898, 0.1, 0.1), (0.1788, 0.1, 0.1), (0.5898, 1, 0.1), (0.1788, 1, 0.1), and (0.1788, 0.1, 0.3). The solid and dashed–dotted lines represent the mass distribution at different times (see Section 3.2). The horizontal axis indicates the moonlet mass (Mp) normalized by MG, and the vertical axis indicates the number of moonlets whose masses are larger than the given moonlet mass. (B) Residence time of a moonlet whose mass is MG formed by streaming instability at r = 3 Rplanet as a function of the planetary mass Mplanet normalized by the Earth mass. The blue solid and yellow dashed–dotted lines represent disks formed by collisions between icy planets and rocky planets, respectively. 8 The Planetary Science Journal, 5:145 (12pp), 2024 June Nakajima et al.
  • 9. could form from an initially vapor-rich disk. Thus, we suggest that an initially vapor-rich disk is not suitable for forming our Moon, and our result supports the hypothesis that the Moon formed from an initially vapor-poor disk, including the canonical model where the proto-Earth was hit by a Mars- sized impactor (Canup & Asphaug 2001). The isotopic composition of the Moon may constrain whether streaming instability played a role during the Moon formation. Some may argue that the observational fact that the Moon is depleted in volatiles may be explained by accreting moonlets formed by streaming instability before volatiles accreted onto the Moon. However, this process would make the Moon isotopically light if these isotopes experience kinetic fractionation (Dauphas et al. 2015, 2022), which is inconsistent with the observation of the enrichment of heavy isotopes in some elements in the Moon (such as K; Wang & Jacobsen 2016). The lunar isotopes would be heavier if they experience equilibrium fractionation, but the equilibrium fractionation would not produce the observed isotopic fractionations under the high disk temperature condition. Alter- natively, some volatiles could be lost from the lunar magma ocean under an equilibrium condition (Charnoz et al. 2021), but the efficiency of the volatile loss is not fully known (Dauphas et al. 2022). 4.2. Streaming Instability in General Impact-induced Disks Our model supports the previous work that suggests that relatively large rocky (>6 M⊕, >1.6 R⊕) and icy (>1 M⊕, > 1.3 R⊕) planets cannot form impact-induced moons that are large compared to the host planets (Nakajima et al. 2022). Larger planets than those thresholds generate completely vapor disks, because the kinetic energy involved in an impact scales with the planetary mass. Thus, these large planets are not capable of forming moons that are large compared to their host planets. Moons can form by mechanisms other than impacts, such as formation in a circumplanetary disk and gravitational capture, but these moons tend to be small compared to the sizes of their host planets (the predicted moon-to-planet mass ratio is ∼10−4 for moons formed by circumplanetary disk, Canup & Ward 2006; and gravitationally captured moons are small in the solar system, Agnor & Hamilton 2006). Thus, fractionally large moons compared to the host planet sizes, which are observationally favorable, likely form by impact. So far, no exomoon has been confirmed despite extensive searches, but future observations, especially with the James Webb Space Telescope (Cassese & Kipping 2022), may be able to find exomoons and test this theoretical hypothesis. 4.3. Comparison between the Moon-forming Disk and Protoplanetary Disk It is certainly intriguing that the streaming instability is able to solve the gas drag problem in the protoplanetary disk but not in the Moon-forming disk. In both scenarios, the clump sizes formed by the streaming instability happen to be similar (MG ∼ 100 km, see Johansen et al. 2015, for the protoplanetary disk and ∼100 km for the protolunar disk, respectively). This size is ∼105 times larger than the size of the particle (0.52–1 m; Adachi et al. 1976) that experiences the strongest gas drag (τf = 1 and vr ∼ −ηvK) in the protoplanetary disk. Therefore, once particles become ∼100 km-sized planetesimals by streaming instability, the radial velocities of the planetesimals decrease drastically (vr v 2 K f = - h t for large τf; see Equation (A7)), which helps planetesimal growth. In contrast, the largest moonlet size (100 km) formed by streaming instability in the Moon- forming disk is only 50 times larger than the particle size that experiences the strongest gas drag (2 km). This does not result in a large τf change or vr change; therefore, streaming instability does not effectively help moon formation. Therefore, streaming instability is an effective mechanism to skip the “1 m barrier” in Figure 4. Schematic view of Moon formation from an initially vapor-rich disk. 9 The Planetary Science Journal, 5:145 (12pp), 2024 June Nakajima et al.
  • 10. the protoplanetary disk, whereas it is not an effective mechanism to skip the “kilometer barrier” in the Moon-forming disk. 4.4. Streaming Instability in the Circumplanetary Disk around a Gas Giant The streaming instability may occur during moon formation in various satellite systems around gas giants. The Galilean moons around Jupiter and Titan around Saturn likely formed from their circumplanetary disks. In these disks, the particle-to-gas ratio Z is typically 10−4 –10−2 , which was thought to be too small for streaming instability to occur (Carrera et al. 2015; Yang et al. 2017; Shibaike et al. 2017). However, recent streaming instability calculations show that streaming instability can occur as low as Z = 4 × 10−3 (at Δ = 0.05 and τf = 0.3; Li & Youdin 2021), and the required Z depends on the disk conditions (e.g., Carrera et al. 2015; Yang et al. 2017; Sekiya & Onishi 2018). This may indicate that streaming instability can be potentially important in circumplanetary disks around gas giants if the dust–gas ratio of the disk is relatively high. The velocity difference between the gas and particles vrel is (Canup & Ward 2002) ⎛ ⎝ ⎞ ⎠ ( ) v v c c r . 12 rel k s s h = ~ W Assuming 0.1 c r c vk s s = ~ W (Canup & Ward 2002), this yields η ∼ 0.01 and 0.1 v c k s D = ~ h . If Z is sufficiently large (at least Z > 4 × 10−3 ; Li & Youdin 2021), it is possible that the streaming instability takes place in a circumplanetary disk around a gas giant. A rough estimate of the size of a moonlet in a circumplanetary disk is MG = 1.4 × 1016 kg (Equation (11)), which is ∼10.3 km in radius if it is a rocky moonlet. Here, we are assuming Z = 10−2 , Σp = 3 × 104 kg m−2 , cs = 1000 m s−1 , H ∼ cs/Ω, and ( ) H 2 g g r p = S (Canup & Ward 2002). These moonlets are relatively small compared to large moons around gas giants (e.g., the Galilean moon masses are in the range of 1022 –1023 kg), and it is unclear if they would significantly impact these moon-formation processes. Further research is needed to understand their effects on the satellite formation in a circumplanetary disk around a gas giant. 4.5. Model Limitations There are several model limitations that need to be addressed in our future work. First, the effect of the Roche radius (aR ∼ 3 R⊕) needs to be taken into account to understand the Moon's formation. An inspiraling moonlet would not directly reach the Earth but would be tidally disrupted near the Roche radius, where it then might be incorporated into an accretion disk (e.g., Salmon & Canup 2014). Second, our model presented here does not take into account the evolution of the disk in detail, which is important to identify the final mass and composition of the resulting moon. As the disk spreads out, it is likely that Δ decreases, which would slow down the radial drift of moonlets. This gas drag effect would disappear once the local vapor condenses, which would occur at the outer part of the disk first, since that part of the disk can cool efficiently due to its large surface area. Given that such moonlets that directly form from the disk would not have time to lose volatiles in the disk phase, this model requires the Moon to have lost its volatiles before or after the disk phase, for example, during the lunar magma ocean phase (Charnoz et al. 2021; see further discussion in Section 4.1). Nevertheless, such a scenario may be possible in moon-forming disks in the solar and extrasolar systems. The effects of the Roche radius and disk evolution will be addressed in our future work. 5. Conclusions In conclusion, we conduct hydrodynamic simulations using Athena and show that the streaming instability can form self- gravitating clumps (∼102 km) in a vapor-rich moon-forming disk generated by a giant impact, but the sizes of the clumps it generates are not large enough to avoid inspiraling due to the strong gas drag. This is a major difference from processes in the protoplanetary disk, where the streaming instability can efficiently form large clumps to avoid the strong gas drag effect. As a result, growing moonlets in an initially vapor-rich moon-forming disk continue to fall onto the planet once they reach sizes of 100 m– 100 km. These moonlets could grow further once the disk cools enough and the VMF of the disk becomes small. However, by this time, a significant amount of the disk mass is lost, and the remaining disk could make only a small moon. This result is applicable to impact-induced moon-forming disks in the solar system and beyond; we find that the streaming instability is not an efficient mechanism to form a large moon from an impact-induced vapor-rich disk in general. As a result, we support previous work that suggests that fractionally large moons compared to their host planets form from vapor-poor disks. This supports the Moon- formation models that produce vapor-poor disks, such as the canonical model. For exomoons, our work supports the previous work that suggests that the ideal planetary radii that host fractionally large moons are 1.3–1.6 R⊕ (Nakajima et al. 2022) given that rocky or icy planets larger than these sizes would likely produce completely vapor disks, which are not capable of forming large moons (Nakajima et al. 2022). The streaming instability may take place in circumplanetary disks, but their effect on the moon- formation process needs further investigation. Acknowledgments We thank the code developers of Athena (Stone et al. 2008; Bai & Stone 2010a; Simon et al. 2016) and PLAN (Li et al. 2019). We appreciate discussions with Shigeru Ida and Scott D. Hull. J.A. was partially supported by the Research Experience for Undergraduates (REU) Program, National Science Founda- tion (NSF), under grant No. PHY-1757062. M.N. was supported in part by the National Aeronautics and Space Administration (NASA) grant Nos. 80NSSC19K0514, 80NSSC21K1184, and 80NSSC22K0107. Partial funding for M.N. was also provided by NSF EAR-2237730 as well as the Center for Matter at Atomic Pressures (CMAP), an NSF Physics Frontier Center, under award PHY-2020249. Any opinions, findings, conclu- sions, or recommendations expressed in this material are those of the authors and do not necessarily reflect those of the National Science Foundation. M.N. was also supported in part by the Alfred P. Sloan Foundation under grant G202114194. Appendix Appendix A The momentum equation for gas in the radial direction is written as ( ) v r GM r dp dr 1 , A1 2 2 g g r = + f Å 10 The Planetary Science Journal, 5:145 (12pp), 2024 June Nakajima et al.
  • 11. where vf is the azimuthal velocity of the gas. Using Equation (2) in the main text and the relationship of p c g g 2 2 r = , ( ) ( ) v v 1 2 . A2 K 1 2 h = - f The equations of motion of the particles in the gas in the radial and azimuthal directions, vr and vf, are (Armitage 2010; Takeuchi & Lin 2002) ( ) ( ) dv dt v r r t v v 1 , A3 r r r 2 2 f ,g = - W - - f ( ) ( ) ( ) d dt rv r t v v , A4 f ,g = - - f f f where vr,g and vf,g are the gas velocities in the radial and azimuthal directions. Assuming that the d dt terms are negligible, one finds that ( ) v v v 2 . A5 r r f 1 ,g K f f 1 t h t t = - + - - This formulation is slightly different from previous formula- tions (−2ηvK instead of −ηvK; Armitage 2010; Takeuchi & Lin 2002) due to the different definition of η by a factor of 2. This leads to ( ) v v v 2 . A6 r r f 1 ,g K f f 1 t h t t = - + - - This can be further simplified by the values of τf as ⎧ ⎨ ⎪ ⎪ ⎩ ⎪ ⎪   ( ) ( ) v v v v v v 2 at 1, 1 2 2 at 1, 2 at 1. A7 r r r K ,g f K f ,g K f f f ht t h t h t t = - - ~ - For a steady disk flow, one could assume vr r ,g 3 2 ~ - n (Armitage 2010). In the main text, we simply assume vr,g = 0. Appendix B An example input file for a 2D Athena simulation is listed in the supplementary material (athinput.par_strat2d). The resolu- tion in this file is 512 × 512, and τf = 1, Z = 0.1, and Δ = 0.1. ORCID iDs Miki Nakajima https:/ /orcid.org/0000-0001-5014-0448 Jeremy Atkins https:/ /orcid.org/0009-0005-0722-8408 Jacob B. Simon https:/ /orcid.org/0000-0002-3771-8054 Alice C. Quillen https:/ /orcid.org/0000-0003-1280-2054 References Abod, C. P., Simon, J. B., Li, R., et al. 2019, AJ, 883, 192 Adachi, I., Hayashi, C., & Nakazawa, K. 1976, PThPh, 56, 1756 Agnor, C. B., & Hamilton, D. P. 2006, Natur, 441, 192 Armitage, P. J. 2010, Astrophysics of Planet Formation (Cambridge: Cambridge Univ. Press) Armytage, R. M. G., Georg, R. B., Williams, H. M., & Halliday, A. N. 2012, GeCoA, 77, 504 Asphaug, E., Emsenhuber, A., Cambioni, S., Gabriel, T. S. J., & Schwartz, S. R. 2021, PSJ, 2, 200 Bai, X.-N., & Stone, J. M. 2010a, ApJ, 722, 1437 Bai, X.-N., & Stone, J. M. 2010b, ApJS, 190, 297 Bonomo, A. S., Zeng, L., Damasso, M., et al. 2019, NatAs, 3, 416 Bouhifd, A. M., Jephcoat, A. P., Porcelli, D., Kelley, S. P., & Marty, B. 2020, GChPL, 15 Cameron, A. G. W., & Ward, W. R. 1976, Lunar Planet. Sci. VII, 120 Cano, E. J., Sharp, Z. D., & Shearer, C. K. 2020, NatGe, 13, 270 Canup, R. M. 2004, Icar, 168, 433 Canup, R. M. 2005, Sci, 307, 546 Canup, R. M. 2012, Sci, 338, 1052 Canup, R. M., & Asphaug, E. 2001, Natur, 412, 708 Canup, R. M., Visscher, C., Salmon, J., & Fegley, B. 2015, NatGe, 8, 918 Canup, R. M., & Ward, W. R. 2002, AJ, 124, 3404 Canup, R. M., & Ward, W. R. 2006, Natur, 441, 834 Canup, R. M., Righter, K., Dauphas, N., et al. 2023, RvMG, 89, 53 Carrera, D., Johansen, A., & Davies, M. B. 2015, A&A, 579, A43 Cassese, B., & Kipping, D. 2022, MNRAS, 516, 3701 Charnoz, S., & Michaut, C. 2015, Icar, 260, 440 Charnoz, S., Sossi, P. A., Lee, Y.-N., et al. 2021, Icar, 364, 114451 Chiang, E., & Youdin, A. N. 2010, AREPS, 38, 493 Colella, P. 1990, JCoPh, 87, 171 Colella, P., & Woodward, P. R. 1984, JCoPh, 54, 174 Craddock, R. A. 2011, Icar, 211, 1150 Cúk, M., Hamilton, D. P., Lock, S. J., & Stewart, S. T. 2016, Natur, 539, 402 Ćuk, M., Lock, S. J., Stewart, S. T., & Hamilton, D. P. 2021, PSJ, 2, 147 Cúk, M., & Stewart, S. T. 2012, Sci, 338, 1047 Dauphas, N. 2017, Natur, 541, 521 Dauphas, N., Poitrasson, F., Burkhardt, C., Kobayashi, H., & Kurosawa, K. 2015, E&PSL, 427, 236 Dauphas, N., Nie, N. X., Blanchard, M., et al. 2022, PSJ, 3, 29 Halliday, A. N., & Canup, R. M. 2022, NRvEE, 4, 19 Halliday, N. H. 2004, Natur, 427, 505 Hartmann, W. K., & Davis, D. R. 1975, Icar, 24, 504 Hosono, N., Karato, S.-i., Makino, J., & Saitoh, T. R. 2019, NatGe, 12, 418 Hull, S. D., Nakajima, M., Hosono, N., Canup, R. M., & Gassmöller, R. 2024, PSJ, 5, 9 Ida, S., Ueta, S., Sasaki, T., & Ishizawa, Y. 2020, NatAs, 4, 880 Johansen, A., Mac Low, M. M., Lacerda, P., & Bizzarro, M. 2015, SciA, 1, e1500109 Johansen, A., Oishi, J. S., Mac Low, M. M., et al. 2007, Natur, 448, 1022 Kegerreis, J. A., Ruiz-Bonilla, S., Eke, V. R., et al. 2022, ApJL, 937, L40 Kenworthy, M., Lock, S., Kennedy, G., et al. 2023, Natur, 622, 251 Kipping, D. M., Forgan, D., Hartman, J., et al. 2013, ApJ, 777, 134 Krapp, L., Benítez-Llambay, P., Gressel, O., & Pessah, M. E. 2019, ApJL, 878, L30 Kruijer, T. S., Archer, G. J., & Kleine, T. 2021, NatGe, 14, 714 Kruijer, T. S., Kleine, T., Fischer-Godde, M., & Sprung, P. 2015, Natur, 520, 534 Li, R., & Youdin, A. N. 2021, ApJ, 919, 107 Li, R., Youdin, A. N., & Simon, J. B. 2019, ApJ, 885, 69 Lock, S. J., Stewart, S. T., Petaev, M. I., et al. 2018, JGRE, 123, 910 Machida, R., & Abe, Y. 2004, ApJ, 617, 633 Melosh, H. J. 2007, M&PS, 42, 2079 Meng, H. Y. A., Su, K. Y. L., Rieke, G. H., et al. 2014, Sci, 345, 1032 Mordasini, C., Alibert, Y., Georgy, C., et al. 2012, A&A, 547, 112 Mullen, P. D., & Gammie, C. F. 2020, ApJL, 903, L15 Nakajima, M., Genda, H., Asphaug, E., & Ida, S. 2022, NatCo, 13, 568 Nakajima, M., & Stevenson, D. J. 2014, Icar, 233, 259 Nakajima, M., & Stevenson, D. J. 2015, E&PSL, 427, 286 Nakajima, M., & Stevenson, D. J. 2018, E&PSL, 487, 117 Nie, N. X., & Dauphas, N. 2019, ApJL, 884, L48 Nielsen, S. G., Bekaert, D. V., & Auro, M. 2021, NatCo, 12, 1817 Pahlevan, K., & Stevenson, D. J. 2007, E&PSL, 262, 438 Rufu, R., Aharonson, O., & Perets, H. B. 2017, NatGe, 10, 89 Rufu, R., & Canup, R. M. 2020, JGRE, 125, e2019JE006312 Salmon, J., & Canup, R. M. 2012, ApJ, 760, 83 Salmon, J., & Canup, R. M. 2014, RSPTA, 372, 20130256 Sekiya, M., & Onishi, I. K. 2018, ApJ, 860, 140 Senft, L. E., & Stewart, S. T. 2008, M&PS, 43, 1993 Shibaike, Y., Okuzumi, S., Sasaki, T., & Ida, S. 2017, ApJ, 846, 81 Simon, J. B., Armitage, P. J., Li, R., & Youdin, A. N. 2016, ApJ, 822, 55 Simon, J. B., Blum, J., Birnstiel, T., & Nesvorný, D. 2022, Comets III, in press (arXiv:2212.04509) Slattery, W. L., Benz, W., & Cameron, A. G. W. 1992, Icar, 99, 167 Sossi, P. A., Moynier, F., & van Zuilen, K. 2018, PNAS, 115, 10920 Stone, J. M., & Gardiner, T. A. 2010, ApJS, 189, 142 Stone, J. M., Gardiner, T. A., Teuben, P., Hawley, J. F., & Simon, J. B. 2008, ApJS, 178, 137 11 The Planetary Science Journal, 5:145 (12pp), 2024 June Nakajima et al.
  • 12. Takahashi, S. Z., & Inutsuka, S.-i. 2014, ApJ, 794, 55 Takeuchi, T., & Lin, D. N. C. 2002, ApJ, 581, 1344 Thiemens, M. M., Sprung, P., Fonseca, R. O. C., Leitzke, F. P., & Münker, C. 2019, NatGe, 12, 696 Thiemens, M. M., Tusch, J., O. C. Fonseca, R., et al. 2021, NatGe, 14, 716 Thompson, C., & Stevenson, D. J. 1988, ApJ, 333, 452 Thompson, M. A., Weinberger, A. J., Keller, L. D., Arnold, J. A., & Stark, C. C. 2019, ApJ, 875, 45 Thompson, S. L., & Lauson, H. S. 1972, Improvements in the Chart D Radiation-Hydrodynamic Code. III: Revised Analytic Equations of State Tech. Rep. SC-RR-71-0714, Sandia National Laboratory Tominaga, R. T., Inutsuka, S.-i., & Takahashi, S. Z. 2018, PASJ, 70, 3 Tominaga, R. T., Takahashi, S. Z., & Inutsuka, S.-I. 2019, ApJ, 881, 53 Toro, E. F. 1999, Riemann Solvers and Numerical Methods for Fluid Dynamics: A Practical Introduction (Dordrecht: Springer) Touboul, M., Puchtel, I. S., & Walker, R. J. 2015, Natur, 520, 530 Wang, K., & Jacobsen, S. B. 2016, Natur, 538, 487 Ward, W. R., Canup, R. M., & Rufu, R. 2020, JGR, 125, e2019JE00626 Weidenschilling, S. J. 1977, MNRAS, 180, 57 Wiechert, U., Halliday, A. N., Lee, D. C., et al. 2001, Sci, 294, 345 Williams, C. D., & Mukhopadhyay, S. 2019, Natur, 565, 78 Williams, C. D., Mukhopadhyay, S., Rudolph, M. L., & Romanowicz, B. 2019, GGG, 20, 4130 Yang, C. C., Johansen, A., & Carrera, D. 2017, A&A, 606, 80 Yang, C.-C., & Zhu, Z. 2021, MNRAS, 508, 5538 Yokochi, R., & Marty, B. 2004, E&PSL, 225, 77 Youdin, A. N. 2011, ApJ, 731, 99 Youdin, A. N., & Goodman, J. 2005, ApJ, 620, 459 Young, E. D., Kohl, I. E., Warren, P. H., et al. 2016, Sci, 351, 493 Zhang, J., Dauphas, N. M. D. A., Leya, I., & Fedkin, A. 2012, NatGe, 1429, 251 12 The Planetary Science Journal, 5:145 (12pp), 2024 June Nakajima et al.