You are on page 1of 12

MNRAS 000, 1–12 (2017) Preprint 13 June 2019 Compiled using MNRAS LATEX style file v3.

Linear Analysis of the Nonaxisymmetric Secular


Gravitational Instability

1⋆
Mohsen Shadmehri , Razieh Oudi 2, Gohar Rastegarzadeh 2
1 Department of Physics, Faculty of Sciences, Golestan University, Gorgan 49138-15739, Iran
2 Department of Physics, Semnan University, Semnan 35196-45399, Iran
arXiv:1906.05256v1 [astro-ph.GA] 12 Jun 2019

Accepted XXX. Received YYY; in original form ZZZ

ABSTRACT
In protoplanetary discs (PPDs) consisting of gas and dust particles, fluid instabilities
induced by the drag force, including secular gravitational instability (SGI) can facili-
tate planet formation. Although SGI subject to the axisymmetric perturbations was
originally studied in the absence of gas feedback and it then generalized using a two-
fluid approach, the fate of the nonaxisymmetric SGI, in either case, is an unexplored
problem. We present a linear perturbation analysis of the nonaxisymmetric SGI in a
PPD by implementing a two-fluid model. We explore the growth of the local, nonax-
isymmetric perturbations using a set of linearized perturbation equations in a sheared
frame. The nonaxisymmetric perturbations display a significant growth during a finite
time interval even when the system is stable against the axisymmetric perturbations.
Furthermore, the surface density perturbations do not show the continuous growth but
are temporally amplified. We also study cases where the dust component undergoes
amplification whereas the gas component remains stable. The amplitude amplification,
however, strongly depends on the model parameters. In the minimum mass solar neb-
ula (MMSN), for instance, the dust fluid amplification at the radial distance 100 au
occurs when the Stokes number is about unity. But the amplification factor reduces as
the dust and gas coupling becomes weaker. Furthermore, perturbations with a larger
azimuthal wavelength exhibit a larger amplification factor.
Key words: accretion – accretion discs – planetary systems: protoplanetary discs

1 INTRODUCTION onto it (Cameron 1973; Hayashi et al. 1977; Lissauer 1993;


D’Angelo et al. 2010).
Understanding planet formation mechanisms in protoplane-
tary discs (PPDs) is still a controversial issue despite consid- Linear perturbation analysis and numerical simulations
erable achievements in recent years (Lissauer & Stevenson show that a gaseous disc with the surface density Σ, Kep-
2007; Durisen et al. 2007; Helled et al. 2014). The most lerian angular velocity Ω and the sound speed cs is grav-
widely studied planet formation theories are the so-called itationally unstable subject to the axisymmetric pertur-
core accretion model which is efficient in the inner re- bations if Toomre parameter, i.e., Q ≡ cs Ω/(πGΣ) be-
gion of a PPD (e.g., Mizuno 1980; Stevenson 1982; comes less than a critical value around unity (Toomre
Pollack et al. 1996) and gravitational instability (GI) which 1964). This criterion has been successfully implemented in
operates mainly in the outer regions (e.g., Adams et al. the star formation theories in galaxies (e.g., Collin & Zahn
1989; Boss 1997; Boley 2009; Rafikov 2009). However, al- 2008; Krumholz & Burkert 2010; Romeo & Agertz 2014;
ternative planet formation theories have been proposed Goldbaum et al. 2016) and planet formation scenarios in
during recent years, including inside-out planet forma- the PPDs (e.g., Matzner & Levin 2005; Kratter et al. 2008;
tion scenario (Chatterjee & Tan 2014) and tidal downsiz- Rafikov 2005; Boley 2009; Rafikov 2009). Gammie (2001)
ing model (Nayakshin 2010). In the core accretion model, suggested that Toomre diagnostic is only a necessary condi-
when the mass of an already formed rocky core becomes tion for disc fragmentation and it does not guarantee to have
about 10M⊕ , this planetary embryo is able to grow further long-lived fragments in a turbulent self-gravitating disc (also
through the accretion of its ambient gas and dust particles see, Rice et al. 2003; Mejı́a et al. 2005; Boss 2017). But re-
cent numerical simulations of the self-gravitating discs have
shown that small enough cooling time scale is not a suffi-
⋆ E-mail: m.shadmehri@gu.ac.ir cient condition for fragmentation (Tsukamoto et al. 2015).

c 2017 The Authors



2 M. Shadmehri, R. Oudi, G. Rastegarzadeh
Takahashi et al. (2016) showed that formation of the spi- of the newly born planets at their early formation phase
ral arms in a PPD and their fragmentation are the essential (Andrews et al. 2011; Mayama et al. 2012; Yen et al. 2016;
phases of a disc fragmentation. They proposed a revised con- Loomis et al. 2017; Hendler et al. 2017; Dipierro et al. 2018;
dition for the fragmentation. van Terwisga et al. 2018). But we note that a planet as-
While the primary focus of most disc stability stud- sociated with these rings or gaps has not yet been ob-
ies is the gas component and its evolution, about one per- served directly. Non-planet-related scenarios, thereby, have
cent of a PPD total mass is attributed to the dust par- also been proposed as mechanisms of the multiple ring
ticles with different sizes (Natta et al. 2007). Presence of formation (e.g., Okuzumi et al. 2016; Suriano et al. 2018).
dust particles provides valuable insights about dynamical Takahashi & Inutsuka (2016) applied their two-fluid model
and chemical structure of a PPD (Vasyunin et al. 2011; of the SGI to explain HL Tau rings resulting from this in-
Akimkin et al. 2013; Woitke et al. 2016; Rab et al. 2017), stability. They studied growth time-scale and unstable wave-
its ionization level (Okuzumi et al. 2011; Akimkin 2015; lengths and proposed that SGI is able to create ring struc-
Ivlev et al. 2016; Rab et al. 2017) and radiative transfer tures in the HL Tau disc. Latter & Rosca (2017), on the
through a disc (Akimkin et al. 2013; Rab et al. 2017). For- other hand, suggested that these ring-like structures prob-
mation of rocky planets or cores of gaseous giant plan- ably unrelated to the SGI. In an alternative mechanism
ets is explained based on the collective instabilities asso- (Okuzumi et al. 2016), however, ring-like feature in the HL
ciated with the dust component. In the early works on the Tau disc is explained by incorporating sintering in a dust
clumping of dust particles either purely dusty discs have growth model.
been considered (Safronov 1972; Goldreich & Ward 1973)
or their implemented approximations were rather restrictive Although the focus of the recent studies is to pro-
(Coradini et al. 1981; Sekiya 1983; Noh et al. 1991). vide an explanation for the ring-like structures, some discs
More recent progress for understanding mechanisms of also exhibit complex non-axisymmetric patterns. Spiral arm-
dust clumping in a PPD relies on the existence of a relative like structures, for instance, have been observed in the
velocity between gas and dust particles and the associated PPDs such as MWC 7588 (Grady et al. 2013; Benisty et al.
drag force. A dust layer is generally believed to be formed 2015) and SAO 206462 (Muto et al. 2012; Garufi et al.
at a PPD midplane due to the sedimentation of these parti- 2013). Spiral arms that result from non-axisymmetric per-
cles. Their rotational velocity is Keplerian, whereas the gas turbations are observed in near infrared scattered light
component is rotating with a sub-Keplerian velocity because (Muto et al. 2012; Wagner et al. 2015) and also in sub-
of the radial gradient of pressure. In a PPD as a mixture of millimeter (Tobin et al. 2016).
gas and dust, therefore, these components may experience a
relative velocity. Since the aerodynamically friction force is These spiral features are commonly studied in terms
directly proportional to this relative velocity, motions of dust of planet-disc interactions (Zhu et al. 2015; Lee 2016;
particles are affected by this force. The gas motion, however, Bae & Zhu 2017). In this scenario, spiral waves are gener-
is unaffected by the dust movement when the dust-to -gas ated due to the interaction of a planer with epicyclic oscil-
density ratio is small. lations in a disc. A set of these waves, which are in phase,
A mechanism of the so-called streaming instability (SI) create constructive interference and can create spiral arms
is triggered because of the dust movement through the (Ogilvie & Lubow 2002; Zhu et al. 2015; Lee 2016).
gas (Youdin & Goodman 2005; Youdin & Lithwick 2007;
Jacquet et al. 2011). However, its efficiency strongly de- In an alternative explanation, however, the spiral arms
pends on the dust-to-gas density ratio and the dust-gas cou- are generated due to the non-axisymmetric development of
pling which is quantified in terms of the Stokes number, the gravitational instability (Rice et al. 2003; Dong et al.
i.e., St = tstop Ω where tstop is the stopping time and Ω is 2015; Tomida et al. 2017; Juhász & Rosotti 2018). Spiral
Keplerian angular velocity. For relatively well-coupled par- arm formation in the galactic scale is a well-studied prob-
ticles (St < 1), a mechanism known as secular gravitational lem known as spiral density wave theory (Lin & Shu 1964;
instability (SGI) is driven by the drag force (Youdin 2011; Julian & Toomre 1966; Toomre 1981), and we think that
Shariff & Cuzzi 2011; Michikoshi et al. 2012). This mecha- a similar approach can be implemented in scenarios of GI-
nism is a dissipative version of the classical Toomre analysis induced structures in PPDs. In this regard, SGI provides a
for the two-component discs (Youdin 2011; Shariff & Cuzzi natural route of ring-like patterns in a PPD. But prior linear
2011; Michikoshi et al. 2012; Takahashi & Inutsuka 2014). studies of SGI are restricted to only axisymmetric pertur-
The SGI has primarily been studied using a single fluid bations. Therefore, the recent observed non-axisymmetric
model where the dust dynamics is treated in a gaseous back- features in PPDs motivates us to investigate SGI subject to
ground with no backreaction of the dust. The onset of SGI, the non-axisymmetric perturbations.
therefore, is found to be unconditional and it is triggered no
matter how dusty layer is thin or thick (Youdin 2011). This In this paper, we generalize SGI to a case with nonax-
interesting feature of the SGI is lost when the gas feedback is isymmetric perturbations in a disc composed of the gas and
included in a two-fluid model (Takahashi & Inutsuka 2014). dust particles. Our sheared two-fluid disc model and the
Numerical simulations of the SGI, however, are needed to main equations are presented in Section 2. We then present
address whether this mechanism leads to an appreciable a linear analysis and a set of ordinary differential equations
enhancement of the dust surface density (Tominaga et al. are obtained for the evolution of the perturbations in Sec-
2018). tion 3. These equations are solved numerically in Section 4.
Recently observed multiple concentric ringlike struc- We conclude with a summary of our main results in Section
tures in the PPDs are commonly interpreted as a sign 5.

MNRAS 000, 1–12 (2017)


Secular Gravitational Instability 3
2 BASIC EQUATIONS where ξ is the dimensionless diffusion coefficient and is de-
fined as
We persist with already studied SGI models and their
general formulation (e.g., Takahashi & Inutsuka 2014), 1 + (St) + 4(St)2
ξ = α[ ]. (7)
however, our focus is to explore fate of the non- (1 + (St)2 )2
axisymmetric perturbations. A shearing sheet model
(Goldreich & Lynden-Bell 1965), as a representation of a ra- The Stokes number depends on the particle size, a, and the
zor thin disc small portion, is constructed where its center mean free path of the molecules, λ. When the size of par-
is at a fixed radial distance r0 . The x-axis is defined in the ticles is smaller than the mean free path of the molecules,
radial direction and the y-axis is oriented in the azimuthal i.e. a < 9λ/4, which is known as Epstein regime, the Stokes
direction. Thus, position of any point in this rotating plane number is written as St = (ρm a/ρg vth )fd−1 Ω, where ρg is
is (x, y) = (r − r0 , r0 (θ − Ωt)), where Ω = (GM/r03 )1/2 is the the gas density and ρm = 2 g cm−3 denotes the homo-
Keplerian angular velocity at r = r0 . Here, M is the star geneous material density of a dust particle. We also have
mass. In our analysis we use the relation Ω = −Ω0 = B − A, vth = (8/π)1/2 cs and fd = [1 + (9π/128)(k∆vk/cs )2 ]1/2 ≃ 1,
where A and B are the standard Oort constants and Ω = Ωz where the relative velocity between dust and gas, i.e. ∆v,
is the Keplerian angular velocity. is much smaller than the sound speed at the disc midplane
Our basic equations, therefore, are the continuity, mo- (e.g., Miyake et al. 2016).
mentum, and the Poisson equations for the dust fluid and
gas component:

3 LINEAR PERTURBATIONS
∂Σg
+ ∇.(Σg Vg ) = 0, (1)
∂t The two-fluid equations admit an equilibrium configuration
with a constant gas surface density Σ0g and a constant dust

∂Vg
 surface density Σ0d . The dust-to-gas density ratio is defined
Σg + (Vg .∇)Vg + 2Ω × Vg − Ω2 r = −c2s ∇Σg via ǫ = Σ0d /Σ0g which is an input model parameter. The
∂t
gas and dust components are undergoing Keplerian motion,
Σd (Vd − Vg )
−Σg ∇(ψg + ψd ) + i.e. V0g = V0d = 2Axj. Therefore, the linearised equations
tstop are
∂ ∂vig ∂vkg 2 ∂vlg
+ [Σg ν( + − δik )], (2) ∂(δΣg ) ∂(δΣg ) ∂(δvxg ) ∂(δvyg )
∂xk ∂xk ∂xi 3 ∂xl + 2Ax + Σ0g ( + ) = 0, (8)
∂t ∂y ∂x ∂y

∂Σd
+ ∇.(Σd Vd ) = D∇2 Σd , (3)
∂t ∂(δvxg ) ∂(δvxg ) ∂
+ 2Ax − 2Ω(δvyg ) = − (δψg + δψd )
∂t ∂y ∂x
c2 ∂(δΣg ) ǫ(δvxd − δvxg )
 
∂Vd − s
Σd + (Vd .∇)Vd + 2Ω × Vd − Ω2 r = −c2d ∇Σd Σ0g ∂x
+
tstop
∂t
Σd (Vg − Vd ) 4 ∂ 2 (δvxg ) ∂ 2 (δvxg ) 2Aν ∂(δΣg ) ∂2
−Σd ∇(ψg + ψd ) + , (4) + ν + ν + + ν (δvyg ),
tstop 3 ∂x2 ∂y 2 Σ0g ∂y ∂y∂x
(9)

∇2 (ψg + ψd ) = 4πG(Σg + Σd )δ(z). (5)


∂(δvyg ) ∂(δvyg ) ∂
The subscripts ”g” and ”d” stand for the gas and dust com- + 2Ax + 2B(δvxg ) = − (δψg + δψd )
∂t ∂y ∂y
ponents. Thus, Σg and Σd are the gas and the dust sur-
c2 ∂(δΣg ) ǫ(δvyd − δvyg )
face densities. The velocities of the gas and the dust fluids − s +
are denoted by Vg and Vd . Furthermore, ψg and ψd rep- Σ0g ∂y tstop
resent the gravitational potential associated with the gas 4 ∂ 2 (δvyg ) ∂ 2 (δvyg ) 2Aν ∂(δΣg ) ∂2
+ ν 2
+ν 2
+ +ν (δvxg ),
and dust components. The gas sound speed and the veloc- 3 ∂y ∂x Σ0g ∂x ∂x∂y
ity dispersion of the dust particles are cs and cd respec- (10)
tively. Here, ν is the kinematic viscosity which is written as
ν = αc2s Ω−1 , where α is the dimensionless measure of the
∂(δΣd ) ∂(δΣd ) ∂(δvxd ) ∂(δvyd )
turbulent strength. Finally, D is the diffusivity of the dust + 2Ax + Σ0d ( + )
∂t ∂y ∂x ∂y
due to the gas turbulence.
In order to close the equations of the system, a relation ∂2 ∂2
= D( 2 + )δΣd , (11)
between cs and cd and an equation for the diffusivity coef- ∂x ∂y 2
ficient D are needed. Youdin & Lithwick (2007) found the
following relation:
∂(δvxd ) ∂(δvxd ) ∂
+ 2Ax − 2Ω(δvyd ) = − (δψg + δψd )
1 + 2(St)2 + 45 (St)3 ∂t ∂y ∂x
c2d = αc2s [ ] (6)
(1 + (St)2 )2 c2 ∂(δΣd ) (δvxg − δvxd )
− d + ,
where St is the Stokes number, i.e. St = tstop Ω0 . Further- Σ0d ∂x tstop
more, the diffusivity of the dust D is written as D = ξc2s Ω−1 (12)
0 ,

MNRAS 000, 1–12 (2017)


4 M. Shadmehri, R. Oudi, G. Rastegarzadeh

∂(δvyd ) ∂(δvyd ) ∂ ∂(δvyd ) c2d


+ 2Ax + 2B(δvxd ) = − (δψg + δψd ) + 2B(δvxd ) = ik y [−(δψg + δψd ) − (δΣd )]
∂t ∂y ∂y ∂t′ Σ0d
c2 ∂(δΣd ) (δvyg − δvyd ) (δvyg − δvyd )
− d + , + , (22)
Σ0d ∂y tstop tstop
(13)
∂2
[−(kx − 2Aky t′ )2 − ky2 + ](δψg + δψd )
∂z ′2
∂2 ∂2 ∂2
( 2
+ 2 + 2 )(δψg + δψd ) = 4πG(δΣg + δΣd )δ(z), (14) = 4πG(δΣg + δΣd )δ(z ′ ). (23)
∂x ∂y ∂z
For perturbations with a non-zero ky , it is more
where δ(z) is Dirac delta function. convenient to re-write equations (17)-(23) in terms of a
For exploring non-axisymmetric perturbations, it is new dimensionless time variable, i.e. τ = 2At′ − kx /ky
more convenient to use a shearing sheet coordinates (Goldreich & Lynden-Bell 1965). Therefore, we obtain
(x′ , y ′ , z ′ ) introduced by Goldreich & Lynden-Bell (1965).
We have ∂(δΣg ) ky
+i Σ0g (δvyg − τ δvxg ) = 0, (24)
∂τ 2A
′ ′ ′ ′
x = x, y = y − 2Axt, z = z, t = t, (15)
∂(δvxg ) Ω ky c2
where − (δvyg ) = −i τ [−(δψg + δψd ) − s (δΣg )]
∂τ A 2A Σ0g
∂ ∂ ∂ ∂ ∂
≡ − 2At′ ′ , ≡ , ǫ(δvxd − δvxg ) 4 ky2 τ 2 ν νky2
∂x ∂x′ ∂y ∂y ∂y ′ + − (δvxg ) − (δvxg )
2Atstop 3 2A 2A
∂ ∂ ∂ ∂ ∂
≡ ′ − 2Ax′ ′ , ≡ . (16) iky ν ky2 ντ
∂t ∂t ∂y ∂z ∂z ′ + (δΣg ) + (δvyg ),
Σ0g 2A
In this sheared system of coordinates, all disc quantities as (25)
perturbed as χ = χ0 +δχ, where δχ is assumed to be small in
comparison to the initial state. The disturbances, denoted by ∂(δvyg ) B ky c2
primes, are proportional to exp[i(kx x′ +ky y ′ )], where kx and + (δvxg ) = i [−(δψg + δψd ) − s (δΣg )]
∂τ A 2A Σ0g
ky are the radial and azimuthal wavenumbers respectively.
Thus, the equations (8)-(14) become ǫ(δvyd − δvyg ) 4 ky2 ν ky2 ντ 2
+ − (δvyg ) − (δvyg )
2Atstop 3 2A 2A
∂(δΣg )
+ Σ0g (ikx − 2iky At′ )δvxg + iky Σ0g (δvyg ) = 0, (17) iky ντ ky2 ντ
∂t′ − (δΣg ) + (δvxg ), (26)
Σ0g 2A

∂(δvxg )
− 2Ω(δvyg ) = (ikx − 2iky At′ )[−(δψg + δψd ) ∂(δΣd ) ky Dky2
∂t′ +i Σ0d (δvyd − τ δvxd ) = − (1 + τ 2 )δΣd , (27)
∂τ 2A 2A
c2 ǫ(δvxd − δvxg ) 4
− s (δΣg )] + − ν(kx − 2ky At′ )2 (δvxg )
Σ0g tstop 3 ∂(δvxd ) Ω ky c2
− (δvyd ) = −i τ [−(δψg + δψd ) − d (δΣd )]
2iAky ν ∂τ A 2A Σ0d
−νky2 (δvxg ) + (δΣg ) − νky (kx − 2ky At′ )(δvyg ),
Σ0g (δvxg − δvxd )
(18) + ,
2Atstop
(28)
∂(δvyg ) c2
+ 2B(δvxg ) = iky [−(δψg + δψd ) − s (δΣg )] ∂(δvyd ) B ky c2
∂t′ Σ0g + (δvxd ) = i [−(δψg + δψd ) − d (δΣd )]
∂τ A 2A Σ0d
ǫ(δvyd − δvyg ) 4 2
+ − νky (δvyg ) − ν(kx − 2ky At′ )2 (δvyg ) (δvyg − δvyd )
tstop 3 + ,
2Atstop
2Aν
+ (ikx − 2iky At′ )(δΣg ) − νky (kx − 2ky At′ )(δvxg ), (29)
Σ0g
(19)
∂2
[−ky2 (1+τ 2 )+ ](δψg +δψd ) = 4πG(δΣg +δΣd )δ(z ′ ). (30)
∂z′2
∂(δΣd ) Upon solving the Poisson equation (30) and using the ap-
+ Σ0d (ikx − 2iky At′ )δvxd + iky Σ0d (δvyd )
∂t′ proximation of the finite thickness of the disc (Vandervoort
= D[−(kx − 2ky At′ )2 − ky2 ]δΣd , (20) 1970; Shu 1984), we obtain the gravitational potential per-
turbation:
!
2πG δΣg
∂(δvxd ) (δψg + δψd ) = − [
− 2Ω(δvyd ) = (ikx − 2iky At′ )[−(δψg + δψd ) 1
ky (1 + τ 2 ) 2
1
1 + ky (1 + τ 2 ) 2 H
∂t′
c2 (δvxg − δvxd ) δΣd
− d (δΣd )] + , (21) + 1 ]. (31)
Σ0d tstop 1 + ky (1 + τ 2 ) 2 Hd

MNRAS 000, 1–12 (2017)


Secular Gravitational Instability 5
p
where H = cs /Ω and Hd = (α/St)H are the gas disc scale
height and the dust scale height respectively. d 2 θg dθg 2τ ǫ αQ2g ηR2 1 + τ4
Equations (24)-(29) and (31) constitute main equations − 2
+ [ 2
− − 2 (
dτ dτ 1 + τ ηSt X (1 + ǫ) 3(1 + τ 2 )
2
of the model to be solved subject to appropriate initial con-
τ2 αQ2g ηR2 Q2g R2
ditions. In the absence of drag force, mathematical forms of + )] + θ g [− τ − (1 + τ 2 )
these equations are similar to Jog (1992) who studied GI in a 1 + τ2 2X 2 (1 + ǫ)2 4X 2 (1 + ǫ)2
two-component disc consisting of the gas and stars to mimic R2 (1 + τ 2 )1/2 dθd ǫ
+ ( )] + ( )
a galaxy. Our second component, however, is the dust fluid X(1 + ǫ) 1 + Qg
(1 + τ 2 )1/2 dτ ηSt
2X(1+ǫ)
and its feedback on the gas is included via the drag force.
Time evolution of the perturbations is studied by solving R2 ǫ (1 + τ 2 )1/2
+θd [ ( p α Qg )
our main equations. For numerical purposes, however, it is X(1 + ǫ) 1 + (1 + τ 2 )1/2
St 2X(1+ǫ)
better to re-write equations (24)-(29) and (31) in terms of
ǫξQ2g R2 Qg R 1
dimensionless variables. In doing so, we introduce the fol- + (1 + τ 2 )] + Wg [ ( )
lowing new variables: 4X 2 St(1 + ǫ)2 X(1 + ǫ) 1 + τ 2
RQg 1 αQ3g ηR3 τ (1 − τ 2 )
− − ]=0
Xη(1 + ǫ) 12 X 3 (1 + ǫ)3 1 + τ 2
(34)

dWd Qg R dθd η − 2
κcs κcd 2A − ( )− ( )
Qg = , Qd = , η= dτ 2X(1 + ǫ) dτ η
πGΣ0g πGΣ0d Ω0 ξQ2g R2 (η − 2) RQg
δvxg δvyg δvxd −θd ( )(1 + τ 2 ) + (Wg − Wd ) = 0
δuxg = , δuyg = , δuxd = , 4X 2 (1 + ǫ)2 2ηStX(1 + ǫ)
cs cs cs (35)
δvyd δΣg δΣd
δuyd = , θg = , θd =
cs Σ0g Σ0d
d 2 θd dθd 2τ 1 ξηQ2g R2
κ2 2(2 − η) λy − + [ − − (1 + τ 2 )]
R2 = = , X= . (32) dτ 2 dτ 1 + τ 2 ηSt 4X 2 (1 + ǫ)2
4A2 η2 λcrit
dθg 1 R2 ǫ (1 + τ 2 )1/2
+ ( ) + θd [ ( p α Qg )
dτ ηSt X(1 + ǫ) 1 + (1 + τ 2 )1/2
St 2X(1+ǫ)

ξQ2g R2 ǫQd R 2
− (1 + τ 2 ) − ( ) (1 + τ 2 )]
4X 2 St(1 + ǫ)2 2X(1 + ǫ)
Here, parameters Qg and Qd stand for the Toomre pa- R2 (1 + τ 2 )1/2
+θg [ ( Qg
)]
rameter for the gas and dust components respectively. The X(1 + ǫ) 1 + (1 + τ 2 )1/2
2X(1+ǫ)
epicyclic frequency is denoted by κ where for a Keplerian Qg R 1 RQg
disc it becomes κ = Ω. Furthermore, the shear parameter +Wd [ ( )− ]=0
X(1 + ǫ) 1 + τ 2 Xη(1 + ǫ)
is η = 2A/Ω0 . Ratios of the gas and dust surface densities (36)
and their corresponding initial values are shown by θg and
θd respectively. We also introduce a critical wavelength, i.e. where Wg = i[δuxg + τ (δuyg )] and Wd = i[δuxd + τ (δuyd )].
λcrit = 4π 2 GΣ0g (1+ǫ)/κ2 , and the perturbation wavelength Equations (33)-(36) are solved using Runge-Kutta method
is written in terms of this critical wavelength. to determine evolution of the perturbations with time. In
Using introduced dimensionless variables, equations the next section, we present our solutions.
(24)-(29) and (31) are reduced to the following set of or-
dinary differential equations:
4 NUMERICAL SOLUTIONS
Although a wide range of the initial conditions can be imple-
mented, we consider the following simple initial conditions.
At the initial time τini , the relevant quantities are
(θg , dθg /dτ, θd , dθd /dτ, Wg , Wd ) = (1, 0, 0, 0, 0, 0). (37)
dθg η − 2 1 αQ2g ηR2 τ (1 − τ 2 ) When a perturbation oscillates with an amplitude less than
− [ + ]
dτ η 6 X 2 (1 + ǫ)2 1 + τ 2 unity, we consider it as a stable configuration. Note that
αQ2g ηR2 dWg Qg R stable axisymmetric perturbations also exhibit oscillatory
−θg ( )(1 − τ 2 ) − ( )
4X 2 (1 + ǫ)2 dτ 2X(1 + ǫ) behavior with time. The non-axisymmetric perturbations,
αQ3g ηR3 however, are unstable once their amplitudes display sig-
RǫQg (1 − τ 2 )2 8τ 2
−Wg [ + 3 3
( 2
+ )] nificant growth only for a limited time. Unstable axisym-
2ηStX(1 + ǫ) 8X (1 + ǫ) 1+τ 3(1 + τ 2 )
metric perturbations grow with an exponential profile, but
RǫQg growth of the nonaxisymmetric perturbations is not expo-
+Wd ( )=0
2ηStX(1 + ǫ) nential and the concept of the instability corresponds to
(33) rapid transient amplification of the perturbations during a

MNRAS 000, 1–12 (2017)


6 M. Shadmehri, R. Oudi, G. Rastegarzadeh

1 st=100 2 st=0.3
dust
dust
1

d
0
d

,
,

g
g

-1 gas -1 gas
0 100 200 300 400 500 -10 0 10 20 30

2 1 st=0.1
st=10 gas

1 dust

d
0
d

,
,

g
0
g

gas -1 dust
-1
0 20 40 60 80 100 -10 0 10 20 30

3 1 gas
st=5 st=0.01
2 dust
0
d
d

1
,
,

-1
g

0
gas dust
-1 -2
0 15 30 -10 0 10 20 30

2
4 dust st=1 gas st=0.001
1
2
d

0
,

,
g

0 -1
-2 gas -2 dust
-10 0 10 20 30 -10 0 10 20 30

Figure 1. The ratio of the surface density perturbation to the unperturbed surface density for the gas (black curve) and dust (red curve)
components, i.e. θg and θd versus the dimensionless time parameter τ subject to the initial conditions (37) with τini = −10. Other model
parameters are Qg = 15, Qd = 11, ǫ = 0.01, X = 3, η = 1.5 and α = 10−4 .

MNRAS 000, 1–12 (2017)


Secular Gravitational Instability 7

2 St=0.3 dust
0

d
,
g
gas
-2
-10 0 10 20 30

2 St=0.7 dust
d

0
, g

gas
-2
-10 0 10 20 30
4
St=1 dust
2 gas
d
,

0
g

-2
-10 0 10 20 30

Figure 2. Evolution of the gas surface density perturbation (black curve) and the dust surface density perturbation (red curve) in MMSN
at the radial distance 100 au and for different Stokes numbers, as labeled. The input parameters are Qg = 17.7, ǫ = 0.01, X = 2.78,
η = 1.5 and α = 10−4 . Also, the dust Toomre parameter is Qd = 18.24, 18.43 and 17.89 corresponding to the Stokes numbers 1, 0.7 and
0.3, respectively.

limited time. These transient amplifications have already Stokes number, as labeled. Obviously, a given Stokes number
been found in the linear stability analysis of the gaseous is equivalent to a certain dust size. Black and red solid curves
self-gravitating discs (Mamatsashvili et al. 2013), discs with correspond to the gas and dust components respectively.
gas and stars (Jog 1992) and even nonaxisymmetric MRI Figure 1 shows that the gas component is always stable
(Balbus & Hawley 1992). It is therefore quite normal that irrespective of the Stokes number. Since the gas Toomre pa-
a mixture of gas and dust undergoes transient amplifica- rameter is larger than its critical value, the gas component
tions subject to the nonaxisymmetric perturbations. The is stable subject to the perturbations and dust dynamics is
fate of these transient patterns can be addressed in the non- unable to change this trend due to a small dust-to-gas den-
linear regime, however, it is important to specify range of sity ratio. The response of the dust component, however,
the model parameters for which the system becomes lin- strongly depends on the adopted Stokes number that con-
early unstable. We thereby consider initial states which are trols the magnitude of the drag force. For a large Stokes
stable subject to the axisymmetric perturbations. But are number, the dust component is also stable because the drag
they remain stable subject to the nonaxisymmetric pertur- force is too weak to affect dust dynamics. But when the
bations? This is an important question that motivated us to Stokes number is less than about 10, the dust component
perform stability analysis with a broad range of the model tends to be unstable subject to the nonaxisymmetric per-
parameters. turbations whereas the gas is still stable. We also find that
We first investigate the evolution of the perturbations the dust component is unstable for St = 0.3. If the Stokes
in Figure 1 for a fiducial set of the model parameters. We number is adopted less than about 0.3, not only the dust
consider an initial configuration with the gas and dust large component becomes stable but also its evolution is similar
Toomre parameters to ensure stability of the system against to the gas evolutionary trend. This behavior is understood
to the axisymmetric perturbations. The model parameters, in terms of the strong dust and gas coupling for the small
therefore, are assumed Qg = 15, Qd = 11, ǫ = 0.01, X = 3, Stokes numbers. Under this condition, the stability of the
η = 1.5 and α = 10−4 . The initial dimensionless time is gas component is dictated to the dust component due to
τini = −10. Note that we also verified that for other values of their strong coupling. Thus, for an intermediate range of the
τini the results are qualitatively similar. We, however, note Stokes number 0.3 . St . 10, while the gas component re-
that the initial time t = 0 corresponds to τini = −kx /ky mains stable, the dust component undergoes transient grow-
that can be rewritten as kx = −τini ky . Since the azimuthal ing patterns during a time interval almost independent of
wavenumber ky is a given model parameter in terms of the the gas component. But either the dust component under-
dimensionless parameter X, we can infer that perturbations goes a growing phase or just display oscillatory behaviour,
with a larger |τini | have a larger radial wavenumber. Each the amplitudes of the perturbations decay at larger times.
panel of Figure 1 shows the evolution of the gas and dust In a realistic case, however, our model parameters can
perturbations, i.e. θg and θd , as a function of τ for a given not be adopted independently as we did in Figure 1. These

MNRAS 000, 1–12 (2017)


8 M. Shadmehri, R. Oudi, G. Rastegarzadeh
ǫ α Qg Qd St a(µm) component remains stable subject to these perturbations ir-
0.01 10−4 17.7 6.66 10 55555 respective of the adopted Stokes number, the dust compo-
0.01 10−4 17.7 18.24 1 5555 nent displays noticeable growth and the amplification factor
0.01 10−4 17.7 17.71 0.1 555 reduces with decreasing the Stokes number from 1 to 0.3.
0.01 10−4 17.7 17.70 0.01 55
0.01 10−4 17.7 17.70 0, 001 5.5
The dust abundance is a key parameter and its role is
0.01 10−4 17.7 17.70 0.0001 0.5 explore in Figure 3 for different values of ǫ, as labeled. As
in the previous figure, the model parameters correspond to
Table 1. Our model parameters in the MMSN model with a solar MMSN model at the radial distance 100 au. We set St = 0.3,
mass host star at the radial distance 100 au. Qg = 17.7, η = 1.5 and α = 10−4 . For ǫ = 0.1, 0.01
and 0.001, we have Qd = 1.78, 17.89 and 178.9 and the
corresponding azimuthal dimensionless parameter becomes
parameters depend on the gaseous disc model and the initial X = 2.5, 2.78 and 2.8, respectively. As before, the inte-
distribution of the dust particles. In agreement with most gration is started at τini = −10 and the initial condition
previous studies in this context, we consider the minimum (37) is implemented to determine evolution of the perturba-
mass solar nebula (MMSN; Hayashi 1981) to represent our tions. The gas component exhibits an oscillatory behavior,
gaseous disc model. However, the entire structure of a PPD however, its amplitude significantly decreases with increas-
is unlikely to be described using the MMSN model and there ing dust abundance. But the dust component undergoes a
are also alternative disc profiles (Nixon et al. 2018). Our sta- slightly higher growth rate when its abundance is increased.
bility analysis, therefore, is done in an MMSN model at a In Figure 4, we investigate the role of the parameters η
given radial distance. In this model, the surface density and and α in the evolution of the disturbances. Model parame-
sound speed are given as power-law functions of the radial ters correspond to MMSN model at the radial distance 100
distance (Hayashi 1981): au and the runs are started at τini = −10. In both pan-
 r − 3 els, we adopt St = 0.3 and ǫ = 0.01. In the top panel, we
Σ(r) = 1.7 × 103
2
g cm−2 , (38) have α = 10−4 and different values of the shear parame-
1au ter are considered, as labeled. For the adopted values of the
shear parameter, i.e. η = 1, 1.5 and 1.8, the associated gas
 r − 1 and dust Toomre parameter become Qg = 249.9, 17.7 and
cs (r) = 1.0 × 105
4
cm s−1 . (39) 2.8, and, Qd = 252.6, 17.89 and 2.83, and, we also have
1au
X = 557.3, 2.78 and 0.07, respectively. In the bottom panel,
In MMSN model with a solar mass host star, the gas Toomre
we assume that Qg = 17.7, X = 2.78 and η = 1.5 and the
parameter can therefore be expressed by
dust Toomre parameter for α = 10−3 , 10−4 and 10−5 be-
 r − 1 comes Qd = 56.57, 17.89 and 5.65, respectively. Note that
4
Qg = 56 . (40) both the gas and dust components are stable subject to the
1au
axisymmetric disturbances. In the top panel of Figure 4, role
Furthermore, the Toomre parameter associated to the dust
of η that quantifies the differential rotation rate is shown for
component becomes
different values of this parameter. All model parameters are
1 similar to the bottom panel and the viscosity coefficient is
56 12  r − 14 1 + 2(St)2 + 5/4(St)3 2

Qd = α (41) α = 10−4 . As the parameter η decreases, the dust compo-
ǫ 1au (1 + (St)2 )2
nent tends to be more unstable. The bottom panel of Figure
The Stokes number therefore at the disc midplane becomes 4 shows that fate of the dust component against to the non-
axisymmetric perturbations strongly depends on the chosen
r  32
 
−7 a
St = 1.8 × 10 . (42) viscosity coefficient. While for α = 10−3 , the dust fluid re-
1µm 1au
mains stable, the amplitude of the perturbations gradually
We can use the above relations for determining model pa- increases as the viscosity coefficient tends to the smaller val-
rameters self-consistently in the MMSN model at a given ues.
radial location. In Table 4, we list our model parameters for In our stability analysis, we introduced the dimension-
the MMSN model at the radial distance 100 au. less parameter X that measures azimuthal wavelength of the
Figure 2 displays the evolution of the gas surface den- perturbations in terms of a critical wavelength λcrit . So far
sity perturbation (black curve) and the dust surface den- this parameter has been included as a fixed value. In Figure
sity (red curve) in MMSN model with M = 1 M⊙ at 5, we explore role of this parameter in evolution of the dust
the radial distance 100 au and for different Stokes num- component in MMSN model at the radial distance 100 au for
bers, as labeled. The gas Toomre parameter is Qg = 17.7, these model parameters: Qg = 17.7, Qd = 21.26, St = 0.3,
whereas the Toomre parameter for the dust component be- ǫ = 0.01, η = 1.5, α = 10−4 . As before, the starting time
comes Qd = 18.24, 18.43 and 17.89 corresponding to the is τini = −10. Each curve is marked with the correspond-
Stokes numbers 1, 0.7 and 0.3, respectively. Note that for ing value of X. Comparison of these curves show that the
these Stokes numbers, dust particles are millimeter-sized in amplitude of the oscillations increases with increasing X.
a range between 1.6 mm and 5.5 mm. We have chosen other In Figure 6, we consider typical cases that satisfy con-
model parameters as ǫ = 0.01, X = 2.78, η = 1.5 and ditions of the axisymmetric stability (Takahashi & Inutsuka
α = 10−4 . We verified that both the gas and dust fluids 2014) but finite amplifications are found subject to the non-
are stable subject to the axisymmetric disturbances. But axisymmetric perturbations. We set X = 2.78, η = 1.5,
the response of the system to the nonaxisymmetric pertur- ǫ = 0.01 and α = 10−4 . In the top left hand panel of Figure
bations is quite different as we see in Figure 2. While the gas 6, variations in θg and θd with τ are shown for St = 0.7,

MNRAS 000, 1–12 (2017)


Secular Gravitational Instability 9

2
gas dust =0.1

g d
0

-10 0 10 20 30
2
dust =0.01
g d

0
gas
-2
-10 0 10 20 30
2
dust =0.001
g d

0
gas
-2
-10 0 10 20 30

Figure 3. The ratio of the perturbation surface density to the unperturbed surface density for the gas (black curve) and dust (red
curve) components, i.e. θg and θd versus the dimensionless time parameter τ for different values of metalicity in MMSN model at the
radial distance 100 au. Evolution of the perturbations is calculated subject to the initial condition (37) with τini = −10. Other model
parameters are St = 0.3, Qg = 17.7, η = 1.5 and α = 10−4 . For ǫ = 0.1, 0.01 and 0.001, we have Qd = 1.78, 17.89 and 212.6 and the
corresponding azimuthal dimensionless parameter becomes X = 2.5, 2.78 and 178.9, respectively.

Qg = 7 and Qd = 7.29. The dust component is weakly wavelength λx = 21.23 au, 4.24 au, 2.12 au and 1.41 au, re-
coupled to the gas component and the corresponding per- spectively. This plot shows that the amplitude of the pertur-
turbations exhibit a significant growth during a finite time bations gradually increases with time, however, its growth
period, whereas the gas component remains stable. In the is suppressed after a certain period of time depending on
top right hand panel, the model parameters are St = 0.9, the radial perturbation wavelength. In the middle panel of
Qg = 7 and Qd = 7.26. The dust component again displays Figure 7, we consider cases with different initial time τini
a fairly strong amplification for this choice of the parame- in an interval from −1 to −20 and the resulting maximum
ters. In the bottom panels, we explore stability of the cases amplitude for the gas and dust components, i.e. (θg )Max
with a smaller Toomre parameter. In the bottom left hand and (θd )Max are shown as a function of the radial pertur-
panel, we set St = 0.7, Qg = 5 and Qd = 5.2, whereas the bation wavenumber λx . The adopted model parameters are
bottom right hand panel corresponds to a case with the same St = 0.7, Qg = 17.7, Qd = 18.43, ǫ = 0.01, X = 2.78,
gas and dust Toomre parameter but with a slightly larger α = 10−4 and for a Keplerian disc with η = 1.5. We find
Stokes number, i.e. St = 0.9. Since the Toomre parameter that for λx ≤ 10 au the dust component generally under-
associated with each component is smaller, the amplitude goes a larger (θd )Max in comparison to the gas component.
of the perturbations are larger for both cases with Stokes In other words, the dust component is more unstable in com-
numbers St = 0.7 and St = 0.9 . We find that amplitude of parison to the gas component for the short radial perturba-
the nonaxisymmetric perturbations increases with time and tion wavelength. In the bottom panel, we display growth
then these oscillations are damping. time (i.e., the time when the amplitude becomes maximum)
In all explored cases so far, we used a fixed initial di- as a function of the radial perturbation wavelength for the
mensionless time τini = −10 which then it corresponds to a explored cases. It shows that both dust and gas component
given radial perturbation wavelength λx if the azimuthal evolves to their maximum amplitude during more or less the
wavelength is treated as a given fixed value. This argu- same time period. This time scale, however, is in an interval
ment is based on an already introduced relation as follows from 1000 yr (for long wavelengths) to 3000 yr (for short
kx = −τini ky . We now investigate evolution of the pertur- wavelengths). We find that evolution of the nonaxisymmet-
bations with different initial dimensionless time τini . This ric growth time scale is relatively fast in comparison to the
analysis, thereby, corresponds to evolution of the pertur- axisymmeric growth time scale.
bations with different radial perturbation wavelength for a
given fixed azimuthal wavelength. In the top panel of Figure
7, we exhibit evolution of the dust component for different
5 CONCLUSIONS
values of the initial time τini , as labeled. The other model
parameters are St = 0.7, ǫ = 0.01, Qg = 17.7, Qd = 18.43, We investigated evolution of the imposed nonaxisymmetric
X = 2.78, η = 1.5 and α = 10−4 . Note that the imple- perturbations in a PPD by treating the system as a mix-
mented values τini = −1, -5, -10 and -15 correspond to radial ture of the coupled gas and dust particles. While response of

MNRAS 000, 1–12 (2017)


10 M. Shadmehri, R. Oudi, G. Rastegarzadeh

4
12 =1
8 =1.8 3 X=7
d

4 =1.5
X=9
0 2

0 30 60 90

d
1
4
=10-5 X=3
0 X=5
2 =10-4
d

0 -1
=10-3
-10 0 10 20 30 40 50
-2
0 20 40

Figure 4. The ratio of the dust surface density perturbation to Figure 5. The ratio of the dust surface density perturbation to
the unperturbed surface density, i.e. θd , versus the dimension- the unperturbed surface density, i.e. θd , versus the dimensionless
less time parameter τ in MMSN model. Top panel is plotted for time parameter τ in the MMSN model and for different values
different values of the shear parameter η, whereas the bottom of X, as labeled. Rest of the model parameters are St = 0.3,
panel corresponds to different values of α. All input parameters Qg = 17.7, Qd = 17.89, ǫ = 0.01, η = 1.5 and α = 10−4 .
are calculated at the radial distance 100 au. In both panels, we
set St = 0.3, ǫ = 0.01 and τini = −10. In the top panel, we have
α = 10−4 and each curve is labeled with the adopted shear pa- - Perturbations with a larger azimuthal wavelength ex-
rameter, i.e. η = 1, 1.5 and 1.8. Corresponding to theses values, hibit a relatively higher amplification.
therefore, we have Qg = 249.9, 17.7 and 2.8, Qd = 252.6, 17.89 - Turbulent coefficient has a stabilizing role in promot-
and 2.83 and X = 557.3, 2.78 and 0.07, respectively. In the bot- ing nonaxisymmetric SGI. In the MMSN model, for example,
tom panel, we have Qg = 17.7, X = 2.78 and η = 1.5 and Toomre we showed that growth of the perturbation is completely
parameter of the dust component for α = 10−3 , 10−4 and 10−5
suppressed when the coefficient is α = 10−3 , whereas for
becomes Qd = 56.57, 17.89 and 5.65 respectively.
smaller values of this coefficient, the amplification factor in-
creases.
The final outcome of the implemented nonaxisymmet-
ric perturbations cannot be addressed using the present
the system subject to the axisymmetric perturbations may linear analysis. Our study demonstrates that nonaxisym-
become unstable with an exponential growth, we find that metric perturbations may lead to dust transient amplifi-
nonaxisymmetric perturbations evolve with an oscillatory cation for a wider range of the model parameters in com-
amplitude. But a strong amplification is found correspond- parison to the axisymmetric perturbations. In the context
ing to the cases where are stable against to the axisymmetric of the nonaxisymmetric gravitational instability in galax-
perturbations. We can now summarize our main results: ies, many authors found similar evolutionary trends in ei-
- Growth of the nonaxisymmetric perturbations is not ther purely gaseous discs or two-fluid (stars and gas) sys-
significant when the Stokes number is very large or very tems (e.g., Julian & Toomre 1966; Jog 1992; Fuchs 2001;
small. But there is always an intermediate range for the Michikoshi & Kokubo 2016; Ghosh & Jog 2018). The im-
Stokes number where a high amplification in dust against to posed nonaxisymmetric perturbations get amplified during
the nonaxisymmetric perturbations is found even when the an initial time interval due to the mutual interplay between
gas component remains stable. However, the upper and lower various physical agents including shear and self-gravity of
limits of the Stokes number corresponding to the instability the disc, but eventually this amplification is suppressed be-
depend on the other model parameter including dust and cause of the disc shear. Our model also predicts a similar
gas Toomre parameters. trend for the evolution of the dust component even when the
- In the MMSN model with a solar mass host star, we gas component remains stable. Although our basic equations
found that amplitude of the nonaxisymmetric perturbations are similar to previous studies relevant to gravitational in-
at the radial distance 100 au increases when the Stokes num- stability in galaxies, there is a significant difference because
ber lies in a range between 10−4 and 10−2 . The amplification each component (gas or dust) is permitted to exchange mo-
factor, however, decreases with increasing the Stokes num- mentum via the drag force. In other words, dust and gas
ber. components are gravitationally coupled and they are sub-

MNRAS 000, 1–12 (2017)


Secular Gravitational Instability 11

3
6
2 ini
ini
ini
4 dust 1

d
dust
4
0
d
-1 ini
d

,
,

2 g
g

2
gas
gas
-10 0 10 20
0

( g)Max,( d)Max
0
gas
dust
0 20 40 0 20 40 2
8

6
dust
6 dust 1
1 10
4
4 x
(AU)
d

d
,

,
g

2 gas
gas

Time/(103yr)
2 gas

dust
0
0
2

0 20 40 60 0 20 40 60
1
1 10
x
(AU)

Figure 6. The ratio of the surface density perturbation to the Figure 7. Top panel shows evolution of dust component pertur-
unperturbed surface density for the gas and dust components, i.e. bations with different initial dimensionless time τini , as labeled.
θg and θd versus the dimensionless time parameter τ in MMSN The other model parameters are St = 0.7, ǫ = 0.01, Qg = 17.7,
model at the radial distance 100 au. As before, the initial condi- Qd = 18.43, X = 2.78, η = 1.5 and α = 10−4 . Middle panel dis-
tion (37) is implemented with τini = −10. In all panels, we have plays the maximum growth amplitude for the gas (black square
X = 2.78, η = 1.5, ǫ = 0.01 and α = 10−4 . But other model symbol) and the dust components (red triangle symbol) as a func-
parameters are St = 0.7, Qg = 7 and Qd = 7.29 (top-left), and tion of the radial perturbation wavelength λx for the model pa-
St = 0.9, Qg = 7 and Qd = 7.26 (top-right), St = 0.7, Qg = 5 rameters same as in figure 2 with St = 0.7. In the bottom panel,
and Qd = 5.2 (bottom-left), St = 0.9, Qg = 5 and Qd = 5.2 the associated growth time (i.e., the time when the amplitude be-
(bottom-right). comes maximum) is shown as a function of the radial perturbation
wavelength.

ACKNOWLEDGEMENTS
We are grateful to referee for a constructive report that
helped us to improve the manuscript. MS is also grateful
ject to the momentum exchange by the drag force. We then to Henrik Latter for his constructive comments.
found that the drag force is able to promote temporal growth
of the dust component even in the cases where both compo-
nents are stable subject to the axisymmetric perturbations. REFERENCES
Furthermore, amplitude amplification of the dust compo-
Adams F. C., Ruden S. P., Shu F. H., 1989, ApJ, 347, 959
nent may persists in configurations where the gaseous disc Akimkin V. V., 2015, Astronomy Reports, 59, 747
responds via oscillations with non growing amplitudes. Akimkin V., Zhukovska S., Wiebe D., Semenov D., Pavlyuchenkov
The above mentioned findings thereby propose that Y., Vasyunin A., Birnstiel T., Henning T., 2013, ApJ, 766, 8
nonaxisymmetric SGI will have a better chance to exist. If Andrews S. M., Wilner D. J., Espaillat C., Hughes A. M., Dulle-
that is the case, does it mean the observed dust rings in mond C. P., McClure M. K., Qi C., Brown J. M., 2011, ApJ,
PPDs are the final outcome of SGI? This important ques- 732, 42
tion can not be adequately addressed with a linear analy- Bae J., Zhu Z., 2017, preprint, (arXiv:1711.08161)
Balbus S. A., Hawley J. F., 1992, ApJ, 400, 610
sis because it is not clear if the resulting nonaxisymmetric
Benisty M., et al., 2015, A&A, 578, L6
perturbation evolves to spiral waves or collapsing fragments.
Boley A. C., 2009, ApJ, 695, L53
But we found that growth time scale of the nonaxisymmetric Boss A. P., 1997, Science, 276, 1836
SGI is very fast. It then implies that the resulting patterns Boss A. P., 2017, ApJ, 836, 53
are less axisymmetric and probably the observed dust rings Cameron A. G. W., 1973, Icarus, 18, 407
are not caused by SGI. Further numerical simulations are Chatterjee S., Tan J. C., 2014, ApJ, 780, 53
needed to address this essential question. Collin S., Zahn J.-P., 2008, A&A, 477, 419

MNRAS 000, 1–12 (2017)


12 M. Shadmehri, R. Oudi, G. Rastegarzadeh
Coradini A., Magni G., Federico C., 1981, A& A, 98, 173 Safronov V. S., 1972, Evolution of the protoplanetary cloud and
D’Angelo G., Durisen R. H., Lissauer J. J., 2010, Giant Planet formation of the earth and planets.
Formation. pp 319–346 Sekiya M., 1983, Progress of Theoretical Physics, 69, 1116
Dipierro G., et al., 2018, MNRAS, 475, 5296 Shariff K., Cuzzi J. N., 2011, ApJ, 738, 73
Dong R., Hall C., Rice K., Chiang E., 2015, ApJ, 812, L32 Shu F. H., 1984, in Greenberg R., Brahic A., eds, IAU Colloq. 75:
Durisen R. H., Boss A. P., Mayer L., Nelson A. F., Quinn T., Rice Planetary Rings. pp 513–561
W. K. M., 2007, Protostars and Planets V, pp 607–622 Stevenson D. J., 1982, Planet. Space Sci., 30, 755
Fuchs B., 2001, A&A, 368, 107 Suriano S. S., Li Z.-Y., Krasnopolsky R., Shang H., 2018,
Gammie C. F., 2001, ApJ, 553, 174 MNRAS, 477, 1239
Garufi A., et al., 2013, A&A, 560, A105 Takahashi S. Z., Inutsuka S.-i., 2014, ApJ, 794, 55
Ghosh S., Jog C. J., 2018, A&A, 617, A47 Takahashi S. Z., Inutsuka S.-i., 2016, AJ, 152, 184
Goldbaum N. J., Krumholz M. R., Forbes J. C., 2016, ApJ, Takahashi S. Z., Tsukamoto Y., Inutsuka S., 2016, MNRAS,
827, 28 458, 3597
Goldreich P., Lynden-Bell D., 1965, MNRAS, 130, 125 Tobin J. J., et al., 2016, Nature, 538, 483
Goldreich P., Ward W. R., 1973, ApJ, 183, 1051 Tomida K., Machida M. N., Hosokawa T., Sakurai Y., Lin C. H.,
Grady C. A., et al., 2013, ApJ, 762, 48 2017, ApJ, 835, L11
Hayashi C., 1981, Progress of Theoretical Physics Supplement, Tominaga R. T., Inutsuka S.-i., Takahashi S. Z., 2018, PASJ, 70, 3
70, 35 Toomre A., 1964, ApJ, 139, 1217
Hayashi C., Nakazawa K., Adachi I., 1977, PASJ, 29, 163 Toomre A., 1981, in Fall S. M., Lynden-Bell D., eds, Structure
Helled R., et al., 2014, Protostars and Planets VI, pp 643–665 and Evolution of Normal Galaxies. pp 111–136
Tsukamoto Y., Takahashi S. Z., Machida M. N., Inutsuka S., 2015,
Hendler N. P., et al., 2017, preprint, (arXiv:1711.09933)
MNRAS, 446, 1175
Ivlev A. V., Akimkin V. V., Caselli P., 2016, ApJ, 833, 92
Vandervoort P. O., 1970, ApJ, 161, 87
Jacquet E., Balbus S., Latter H., 2011, MNRAS, 415, 3591
Vasyunin A. I., Wiebe D. S., Birnstiel T., Zhukovska S., Henning
Jog C. J., 1992, ApJ, 390, 378
T., Dullemond C. P., 2011, ApJ, 727, 76
Juhász A., Rosotti G. P., 2018, MNRAS, 474, L32
Wagner K., Apai D., Kasper M., Robberto M., 2015, ApJ, 813, L2
Julian W. H., Toomre A., 1966, ApJ, 146, 810
Woitke P., et al., 2016, A&A, 586, A103
Kratter K. M., Matzner C. D., Krumholz M. R., 2008, ApJ,
Yen H.-W., Liu H. B., Gu P.-G., Hirano N., Lee C.-F., Puspitan-
681, 375
ingrum E., Takakuwa S., 2016, ApJ, 820, L25
Krumholz M., Burkert A., 2010, ApJ, 724, 895
Youdin A. N., 2011, ApJ, 731, 99
Latter H. N., Rosca R., 2017, MNRAS, 464, 1923 Youdin A. N., Goodman J., 2005, ApJ, 620, 459
Lee W.-K., 2016, ApJ, 832, 166 Youdin A. N., Lithwick Y., 2007, Icarus, 192, 588
Lin C. C., Shu F. H., 1964, ApJ, 140, 646 Zhu Z., Dong R., Stone J. M., Rafikov R. R., 2015, ApJ, 813, 88
Lissauer J. J., 1993, ARA&A, 31, 129 van Terwisga S. E., et al., 2018, preprint, (arXiv:1805.03221)
Lissauer J. J., Stevenson D. J., 2007, Protostars and Planets V,
pp 591–606
Loomis R. A., Öberg K. I., Andrews S. M., MacGregor M. A., This paper has been typeset from a TEX/LATEX file prepared by
2017, ApJ, 840, 23 the author.
Mamatsashvili G. R., Chagelishvili G. D., Bodo G., Rossi P., 2013,
MNRAS, 435, 2552
Matzner C. D., Levin Y., 2005, ApJ, 628, 817
Mayama S., et al., 2012, ApJ, 760, L26
Mejı́a A. C., Durisen R. H., Pickett M. K., Cai K., 2005, ApJ,
619, 1098
Michikoshi S., Kokubo E., 2016, ApJ, 823, 121
Michikoshi S., Kokubo E., Inutsuka S.-i., 2012, ApJ, 746, 35
Miyake T., Suzuki T. K., Inutsuka S.-i., 2016, ApJ, 821, 3
Mizuno H., 1980, Progress of Theoretical Physics, 64, 544
Muto T., et al., 2012, ApJ, 748, L22
Natta A., Testi L., Calvet N., Henning T., Waters R., Wilner D.,
2007, Protostars and Planets V, pp 767–781
Nayakshin S., 2010, MNRAS, 408, L36
Nixon C. J., King A. R., Pringle J. E., 2018, MNRAS, 477, 3273
Noh H., Vishniac E. T., Cochran W. D., 1991, ApJ, 383, 372
Ogilvie G. I., Lubow S. H., 2002, MNRAS, 330, 950
Okuzumi S., Tanaka H., Takeuchi T., Sakagami M.-a., 2011, ApJ,
731, 96
Okuzumi S., Momose M., Sirono S.-i., Kobayashi H., Tanaka H.,
2016, ApJ, 821, 82
Pollack J. B., Hubickyj O., Bodenheimer P., Lissauer J. J.,
Podolak M., Greenzweig Y., 1996, Icarus, 124, 62
Rab C., Güdel M., Woitke P., Kamp I., Thi W.-F., Min M., Aresu
G., Meijerink R., 2017, preprint, (arXiv:1711.07249)
Rafikov R. R., 2005, ApJ, 621, L69
Rafikov R. R., 2009, ApJ, 704, 281
Rice W. K. M., Armitage P. J., Bate M. R., Bonnell I. A., 2003,
MNRAS, 339, 1025
Romeo A. B., Agertz O., 2014, MNRAS, 442, 1230

MNRAS 000, 1–12 (2017)

You might also like