Dynamics of Porous Dust Aggregates and Gravitational Instability of Their Disk

and

Published 2017 June 14 © 2017. The American Astronomical Society. All rights reserved.
, , Citation Shugo Michikoshi and Eiichiro Kokubo 2017 ApJ 842 61 DOI 10.3847/1538-4357/aa7388

Download Article PDF
DownloadArticle ePub

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

0004-637X/842/1/61

Abstract

We consider the dynamics of porous icy dust aggregates in a turbulent gas disk and investigate the stability of the disk. We evaluate the random velocity of porous dust aggregates by considering their self-gravity, collisions, aerodynamic drag, turbulent stirring, and scattering due to gas. We extend our previous work by introducing the anisotropic velocity dispersion and the relaxation time of the random velocity. We find the minimum mass solar nebula model to be gravitationally unstable if the turbulent viscosity parameter α is less than about $4\times {10}^{-3}$. The upper limit of α for the onset of gravitational instability is derived as a function of the disk parameters. We discuss the implications of the gravitational instability for planetesimal formation.

Export citation and abstract BibTeX RIS

1. Introduction

Planetesimals are building blocks of planets (e.g., Safronov 1971; Hayashi et al. 1985). The terrestrial planets and the cores of gas giants are considered to be formed by the collisional accretion of planetesimals, which is the so-called core accretion scenario (e.g., Kokubo & Ida 2012). Recently, another scenario, called the pebble accretion, has been proposed (Lambrechts & Johansen 2012, 2014). The centimeter-sized pebbles, loosely coupled with gas, accrete efficiently onto planetesimals owing to gas drag, which leads to the rapid growth of gas giant cores compared to the core accretion scenario. It is not yet understood how dust grains grow into planetesimals, because this process must overcome various obstacles. One such obstacle is the rapid radial drift of dust particles. Because of the radial gas pressure gradient, the gas rotational velocity is slightly slower than the Keplerian velocity; however, dust particles tend to rotate with the Keplerian velocity. Thus, dust particles experience a headwind and lose their angular momentum, and this causes an inward radial drift (Adachi et al. 1976; Weidenschilling 1977b). For example, the meter-sized dust particles fall into the central star in ${10}^{2}\mbox{--}{10}^{3}$ yr before they grow into kilometer-sized planetesimals. The meter-sized dust particles fall into the central star before they grow into kilometer-sized planetesimals.

The gravitational instability (GI) model provides a possible solution to this problem (Safronov 1972; Goldreich & Ward 1973). If the gas flow is laminar, dust particles settle onto the disk midplane, due to the gravitational force of the central star. As a result, the dust layer becomes very thin. When the dust layer density exceeds the Roche density, it becomes gravitationally unstable (Sekiya 1983; Yamoto & Sekiya 2004, 2006). Consequently, dust-rich gas clumps are formed on the dynamical timescale, and planetesimals form in these clumps (Sekiya 1983; Cuzzi et al. 2008). The timescale of the dust-rich gas clump formation is much faster than that of the radial drift, and so the GI model was considered to be a promising mechanism for planetesimal formation.

However, the turbulence in the gas stirs dust particles and prevents them from settling (Weidenschilling 1980; Youdin & Lithwick 2007), and even weak turbulence is sufficient to inhibit the GI. There are various mechanisms that induce turbulence. If the gas is sufficiently ionized, then magnetorotational instability causes strong turbulence (Balbus & Hawley 1991; Sano et al. 2000). Even if the gas flow is initially laminar, sedimentation of dust will cause vertical shear instability, which leads to turbulence (Weidenschilling 1980; Sekiya & Ishitsu 2000, 2001; Ishitsu & Sekiya 2002, 2003; Garaud & Lin 2004; Michikoshi & Inutsuka 2006), and in the minimum mass solar nebula (MMSN) model this instability suppresses the GI (Sekiya 1998). Thus, the classic GI model is unlikely to fulfill its promise for explaining planetesimal formation.

Currently, the classic GI model has been replaced with various other proposed models. One approach considers the streaming instability that is caused by the interaction between gas and dust (Youdin & Goodman 2005). During the nonlinear stage, this instability causes spontaneous particle clumping, which leads to the formation of gravitationally bound objects (Johansen et al. 2007, 2009; Bai & Stone 2010a, 2010b). This clumping is effective if the Stokes number is close to unity and the dust-to-gas mass ratio is large (Carrera et al. 2015; Yang et al. 2016). Several mechanisms that increase the dust-to-gas ratio have been proposed; these include the inward radial drift of dust (Youdin & Shu 2002), a migration trap in the pressure bump (Haghighipour & Boss 2003; Kato et al. 2012; Taki et al. 2016), and the disk wind (Suzuki & Inutsuka 2009; Bai & Stone 2013). In addition, the secular GI forms a dust-rich ring, which may result in planetesimal formation (Youdin 2011; Michikoshi et al. 2012; Takahashi & Inutsuka 2014; Shadmehri 2016; Latter & Rosca 2017).

Another approach considers the pairwise coagulation of fluffy or porous dust aggregates. Studies on dust growth have shown that the icy dust aggregates formed by pairwise coagulation can be significantly porous (Dominik & Tielens 1997; Blum & Wurm 2000; Wada et al. 2007, 2008, 2009; Suyama et al. 2008, 2012); their density is typically ${10}^{-5}\,{\rm{g}}\,{\mathrm{cm}}^{-3}$, which is much smaller than that of a compact object, for which a typical density is $\sim 1\,{\rm{g}}\,{\mathrm{cm}}^{-3}$. In the pairwise coagulation model, the main obstacle is the fragmentation barrier (Blum & Wurm 2008). The critical velocity of fragmentation for porous icy dust aggregates is larger than that for porous silicate dust aggregates (Blum & Wurm 2008; Wada et al. 2009), and so the fragmentation barrier can be overcome relatively easily for icy dust. Thus, in this paper we focus on the evolution of icy dust aggregates. We note that Arakawa & Nakamoto (2016) pointed out that porous silicate dust aggregates consisting of nanometer-sized grains can overcome the fragmentation barrier.

A study of the evolution of porous icy dust aggregates showed that beyond the radial drift barrier, they can grow by pairwise coagulation (Okuzumi et al. 2009, 2012). The timescale of the dust growth and that of the streaming instability are comparable when the Stokes number is unity (Okuzumi et al. 2012). It is not well understood which mechanism is dominant. If the streaming instability works effectively and the Stokes number is close to unity, then gravitationally bound objects form. On the other hand, if pairwise coagulation is more effective than the streaming instability, then the dust aggregates grow and the Stokes number exceeds unity, and this suppresses the streaming instability. In this paper, we postulate that dust aggregates grow sufficiently massive by pairwise coagulation and the Stokes number exceeds unity.

If large dust aggregates form by pairwise coagulation, their density is much smaller than that of compact planetesimals, and so it is clear that some compression mechanism is necessary. Kataoka et al. (2013a) investigated the compression strength of icy dust aggregates; Kataoka et al. (2013b) used these results to evaluate the gas pressure compression and the self-gravity compression and thus track the evolution from porous dust aggregates to compact planetesimals. They found that dust aggregates with mass ${m}_{{\rm{d}}}\gtrsim {10}^{11}\,{\rm{g}}$ are compressed by self-gravity, and the density for a mass of ${10}^{18}\,{\rm{g}}$ reaches $0.1\,{\rm{g}}\,{\mathrm{cm}}^{-3}$. They concluded that planetesimals can form only by pairwise coagulation.

In Michikoshi & Kokubo (2016c, hereafter Paper I), we considered the final stage of the evolution of icy dust aggregates. We investigated the dynamics of dust aggregates and obtained their random velocity. We found that, for a reasonable range of turbulence strength, the porous dust disk becomes gravitationally unstable as the dust aggregates evolve through self-gravity compression. In Paper I, we adopted the MMSN model, and for simplicity we assumed an isotropic velocity dispersion and an equilibrium random velocity. In this paper, we adopt a more general disk model and a more precise dynamical model and confirm the results of Paper I. Here, we consider an anisotropic velocity dispersion, that is, the evolution of the eccentricity and that of the inclination are calculated separately. The relaxation time of the random velocity is also taken into account, since in some cases it is comparable to the timescale of dust growth.

The remainder of the paper is organized as follows. In Section 2, we present the model of the protoplanetary disk and dust aggregates. In Section 3, we explain the porous dust dynamics model. In Section 4, we calculate the equilibrium eccentricity and inclination and investigate the stability of the porous dust disk. We discuss the general properties of the disk independently from the evolution of the dust. In Section 5, we investigate the disk stability as it undergoes compression by self-gravity, and we derive a condition for the onset of GI. Section 6 is devoted to a summary and discussion of our results.

2. Model

2.1. Protoplanetary Disk

We adopt the following surface densities of gas and dust (Weidenschilling 1977a; Hayashi 1981):

Equation (1)

Equation (2)

where a is the distance from the central star, ${\beta }_{{\rm{g}}}$ is the power-law index, fg is the ratio in the MMSN model, and γ is the dust-to-gas mass ratio. As the fiducial model, we consider the MMSN model (${\beta }_{{\rm{g}}}=3/2$ and $\gamma =0.018$) beyond the snowline (Hayashi 1981; Hayashi et al. 1985). We adopt the temperature profile

Equation (3)

where T1 is the temperature at $1\,\mathrm{au}$ and ${\beta }_{{\rm{t}}}$ is the power-law index. In the fiducial model, we adopt ${T}_{1}=120$ and ${\beta }_{{\rm{t}}}=3/7$ (Chiang & Youdin 2010). The isothermal sound velocity is calculated from the temperature as ${c}_{{\rm{s}}}=\sqrt{{k}_{{\rm{B}}}T/{m}_{{\rm{g}}}}$, where kB is the Boltzmann constant and ${m}_{{\rm{g}}}=3.9\times {10}^{-24}\,{\rm{g}}$ is the mean molecular mass. The gas density at the disk midplane is ${\rho }_{{\rm{g}}}={{\rm{\Sigma }}}_{{\rm{g}}}/(\sqrt{2\pi }{c}_{{\rm{s}}}/{\rm{\Omega }})$, where ${\rm{\Omega }}=\sqrt{{{GM}}_{* }/{a}^{3}}$ is the Keplerian angular frequency and ${M}_{* }$ is the mass of the central star. Throughout this paper, we adopt ${M}_{* }={M}_{\odot }$, where ${M}_{\odot }$ is the solar mass. The mean free path of gas molecules is $l={m}_{{\rm{g}}}/{\sigma }_{{\rm{g}}}{\rho }_{{\rm{g}}}$, where ${\sigma }_{{\rm{g}}}=2\times {10}^{-15}\,{\mathrm{cm}}^{2}$ is the collisional cross section of gas molecules. The nondimensional radial pressure gradient is given as (Nakagawa et al. 1986)

Equation (4)

For the fiducial model (${T}_{1}=120$, ${\beta }_{{\rm{g}}}=3/2$ and ${\beta }_{{\rm{t}}}=3/7$), η at 1 au is $0.77\times {10}^{-3}$.

2.2. Dust Aggregate

2.2.1. Physical Properties

For simplicity, we will consider equal-mass dust aggregates with mass md. We assume that a dust aggregate consists of N monomers, where $N={m}_{{\rm{d}}}/{m}_{0}$. The monomer mass, density, and radius are m0, ${\rho }_{0}$, and r0, respectively, which satisfy ${m}_{0}=4\pi /3{\rho }_{0}{r}_{0}^{3}$. The gyration radius of a dust aggregate is given as (Mukai et al. 1992; Wada et al. 2008)

Equation (5)

where ${{\boldsymbol{x}}}_{k}$ is the position of monomer k and ${\boldsymbol{X}}={\sum }_{k}{{\boldsymbol{x}}}_{k}/N$ is the position of the center of mass. The characteristic radius is defined by (Mukai et al. 1992)

Equation (6)

If the dust aggregates grow by ballistic cluster–particle aggregation (BCPA), each dust aggregate can be considered as a uniform sphere, where rc corresponds to the physical radius. On the other hand, if the dust aggregates grow by ballistic cluster–cluster aggregation (BCCA), the dust aggregates are inhomogeneous with a fractal dimension of less than 3, and rc corresponds to the maximum distance from the approximate center of mass (Okuzumi et al. 2009). We define the collisional cross section by using rc. The characteristic volume and the mean internal density are calculated from rc as follows:

Equation (7)

Equation (8)

The density of a dust aggregate with fractal dimension D is

Equation (9)

The fractal dimension of a BCCA cluster is $D\simeq 1.9$ (Mukai et al. 1992; Okuzumi et al. 2009), and thus the density of a BCCA cluster is ${\rho }_{\mathrm{BCCA}}={({m}_{{\rm{d}}}/{m}_{0})}^{-0.58}{\rho }_{0}$.

The interaction of dust aggregates with gas is characterized by the projected area A. Generally, the projected area A can be smaller than $\pi {r}_{{\rm{c}}}^{2}$. Okuzumi et al. (2009) obtained the projected area formula for both BCPA and BCCA clusters

Equation (10)

where ABCCA is the projected area for the BCCA cluster (Minato et al. 2006)

Equation (11)

and ${r}_{{\rm{c}},\mathrm{BCCA}}$ is the characteristic radius of the corresponding BCCA cluster,

Equation (12)

We define the area-equivalent radius from A as

Equation (13)

The area-equivalent radius is shown in Figure 1. If the density is sufficiently larger than the BCCA cluster density ${\rho }_{\mathrm{BCCA}}$, rA is approximated by rc. Since we investigate the evolution due to self-gravity compression, we consider the following range of parameters: ${m}_{{\rm{d}}}\gt {10}^{10}\,{\rm{g}}$ and ${\rho }_{\mathrm{int}}\gt {10}^{-5}\,{\rm{g}}\,{\mathrm{cm}}^{-3}$ (Kataoka et al. 2013b). In this parameter region, the density is much larger than ${\rho }_{\mathrm{BCCA}}$, and the corresponding fractal dimension is larger than 2. Thus, throughout this paper we assume ${r}_{{\rm{A}}}\simeq {r}_{{\rm{c}}}$. In short, we consider a dust aggregate to be a sphere with radius rc, which has the usual mass–radius relation ${m}_{{\rm{d}}}=(4\pi /3){\rho }_{\mathrm{int}}{r}_{{\rm{c}}}^{3}$.

Figure 1.

Figure 1. Area-equivalent radius normalized by the characteristic radius ${r}_{{\rm{A}}}/{r}_{{\rm{c}}}$ on the md${\rho }_{\mathrm{int}}$ plane. The dashed and dotted lines show ${r}_{{\rm{A}}}/{r}_{{\rm{c}}}=0.999$ and 0.99, respectively. The solid line shows the BCCA cluster density ${\rho }_{\mathrm{BCCA}}$.

Standard image High-resolution image

2.2.2. Evolution

We consider the evolution of dust due to coagulation and self-gravity compression (Kataoka et al. 2013b). The compressive strength of a porous dust aggregate is (Kataoka et al. 2013a)

Equation (14)

where Eroll is the rolling energy (Dominik & Tielens 1997; Wada et al. 2007). Equilibrium is obtained when ${P}_{\mathrm{comp}}=P$ (Kataoka et al. 2013b), and from this we can calculate the equilibrium density. The self-gravity pressure is

Equation (15)

and thus we obtain the equilibrium density (Kataoka et al. 2013b)

Equation (16)

As the mass of the dust aggregate increases, its self-gravity pressure increases; this compresses it and increases its equilibrium density. In the fiducial model, we adopt ${E}_{\mathrm{roll}}=4.74\times {10}^{-9}\,\mathrm{erg}$, ${\rho }_{0}=1.0\,{\rm{g}}\,{\mathrm{cm}}^{-3}$, and ${r}_{0}=0.1\,\mu {\rm{m}}$.

3. Dynamics

We define ${\sigma }_{e}$ and ${\sigma }_{i}$ as the rms of the eccentricity and inclination, respectively, of the dust aggregates. We developed a model to calculate the evolution of ${\sigma }_{e}$ and ${\sigma }_{i}$, taking into account gravitational scattering, collisions among dust aggregates, and interactions with gas.

3.1. Gravitational Scattering

Gravitational scattering among dust aggregates results in an increase in both ${\sigma }_{e}$ and ${\sigma }_{i}$ (Ida 1990; Stewart & Ida 2000). In Paper I, we adopted the simple stirring rate formula described by the Chandrasekhar relaxation time (Ida 1990), which is valid in the dispersion-dominated regime. Ohtsuki et al. (2002) examined the evolution of ${\sigma }_{e}$ and ${\sigma }_{i}$ using three-body integration, and they derived a semianalytic formula that is valid in both the dispersion-dominated and the shear-dominated regime. In this paper, we adopt their stirring rates:

Equation (17)

and

Equation (18)

where h is the reduced Hill radius $h={(2{m}_{{\rm{d}}}/3{M}_{* })}^{1/3}$ and

Equation (19)

Equation (20)

where $\beta ={\tilde{\sigma }}_{i,{\rm{r}}}/{\tilde{\sigma }}_{e,{\rm{r}}}$, ${K}_{\lambda }=K(\sqrt{3(1-{\lambda }^{2})}/2)$, ${E}_{\lambda }\,=E(\sqrt{3(1-{\lambda }^{2})}/2)$, and ${\rm{\Lambda }}={\tilde{\sigma }}_{i,{\rm{r}}}({\tilde{\sigma }}_{e,{\rm{r}}}^{2}+{\tilde{\sigma }}_{i,{\rm{r}}}^{2})/12$. The first and second terms correspond to the low-velocity and high-velocity limits, respectively. The reduced relative eccentricity ${\tilde{\sigma }}_{e,{\rm{r}}}$ and the inclination ${\tilde{\sigma }}_{i,{\rm{r}}}$ are ${\tilde{\sigma }}_{e,{\rm{r}}}=\sqrt{2}{\sigma }_{e}/h$ and ${\tilde{\sigma }}_{i,{\rm{r}}}=\sqrt{2}{\sigma }_{i}/h$, respectively (Nakazawa et al. 1989). The functions $K{(k)={\int }_{0}^{\pi /2}(1-{k}^{2}{\sin }^{2}\theta )}^{-1/2}d\theta $ and $E{(k)={\int }_{0}^{\pi /2}(1-{k}^{2}{\sin }^{2}\theta )}^{1/2}d\theta $ are the complete elliptic integrals of the first and second kind, respectively, where k is the modulus.

3.2. Collisions

We assume that the average change in ${\sigma }_{e}$ for a dust collision is ${\sigma }_{e}^{2}\to {C}_{\mathrm{col}}{\sigma }_{e}^{2}$. We adopt ${C}_{\mathrm{col}}=1/2$, which corresponds to perfect accretion (Inaba et al. 2001). We discuss the effect of the imperfect accretion on the dust aggregate growth in Section 5.2. Using the nondimensional collision rate Pcol (Nakazawa et al. 1989), the evolution equations for ${\sigma }_{e}$ and ${\sigma }_{i}$ are written as

Equation (21)

Equation (22)

For the low-velocity regime, where ${\tilde{\sigma }}_{e},{\tilde{\sigma }}_{i}\lt 0.2$, Pcol is independent of ${\tilde{\sigma }}_{e}$ and ${\tilde{\sigma }}_{i}$ (Ida & Nakazawa 1989; Inaba et al. 2001), and

Equation (23)

where ${\tilde{\sigma }}_{e}={\sigma }_{e}/h$ and ${\tilde{\sigma }}_{i}={\sigma }_{i}/h$ are the reduced eccentricity and inclination, respectively, and $\tilde{r}=2{r}_{{\rm{c}}}/{ha}$. For the medium-velocity regime, where $0.2\lt {\tilde{\sigma }}_{e},{\tilde{\sigma }}_{i}\lt 2$ (Ida & Nakazawa 1989; Inaba et al. 2001), Pcol depends on ${\tilde{\sigma }}_{i,{\rm{r}}}$:

Equation (24)

For the high-velocity regime, where $2\lt {\tilde{\sigma }}_{e},{\tilde{\sigma }}_{i}$, Pcol is (Greenzweig & Lissauer 1992)

Equation (25)

where the second term indicates the effect of gravitational focusing. Inaba et al. (2001) proposed the following formula for the general nondimensional collision rate:

Equation (26)

We set ${P}_{\mathrm{col}}={P}_{\mathrm{high}}$ for a large random velocity, such as ${\tilde{\sigma }}_{i}\gt 10$. Note that this formula does not take into account the gas drag effect, which may alter the collision rate. In the present model, when the gas drag is strong, turbulent stirring causes the random velocity to be high. In this case, the collision rate formula is described by the geometrical cross section.

3.3. Gas Effects

3.3.1. Gas Drag

Because of the hydrodynamic gas drag, ${\sigma }_{e}$ and ${\sigma }_{i}$ decrease with time as follows (Adachi et al. 1976; Inaba et al. 2001):

Equation (27)

Equation (28)

where ${t}_{{\rm{s}},e}$ and ${t}_{{\rm{s}},i}$ are the damping timescales of ${\sigma }_{e}$ and ${\sigma }_{i}$,

Equation (29)

Equation (30)

where $E=E(\sqrt{3/4})$ is the elliptic integral of the second kind of argument $\sqrt{3/4}$, ${v}_{{\rm{K}}}=a{\rm{\Omega }}$ is the Keplerian velocity, and CD is the nondimensional drag coefficient. We also adopted the corrections discussed by Kary et al. (1993) and Inaba et al. (2001).

The relative velocity between gas and dust is (Adachi et al. 1976)

Equation (31)

where ${v}_{\mathrm{ran}}={\left(\tfrac{5}{8}{\sigma }_{e}^{2}+\tfrac{1}{2}{\sigma }_{i}^{2}\right)}^{1/2}{v}_{{\rm{K}}}$ is the random velocity. Using the relative velocity given by Equation (31), we define the stopping time

Equation (32)

which is nearly equal to ${t}_{{\rm{s}},e}$ and ${t}_{{\rm{s}},i}$.

The gas drag law changes with rc (e.g., Adachi et al. 1976). If ${r}_{{\rm{c}}}\gtrsim l$, we use the Stokes drag or the Newton drag. For a low Reynolds number ($\mathrm{Re}\,\ll \,{10}^{3}$), the drag coefficient is approximated by ${C}_{{\rm{D}}}\simeq 24/\mathrm{Re}$ (Stokes drag), where $\mathrm{Re}\,=\,2{r}_{{\rm{c}}}u/\nu $. The viscosity ν is given by $\nu ={v}_{\mathrm{th}}l/2$, where ${v}_{\mathrm{th}}=\sqrt{8/\pi }{c}_{{\rm{s}}}$ is the thermal velocity. For a high Reynolds number (${10}^{3}\lt \mathrm{Re}\,\lt \,2\times {10}^{5}$), the drag coefficient is almost constant, ${C}_{{\rm{D}}}\simeq 0.4\mbox{--}0.5$ (Newton drag). If ${r}_{{\rm{c}}}\lesssim l$, we use the Epstein drag. Thus, we adopt the drag coefficient formula (Brown & Lawler 2003)

Equation (33)

3.3.2. Turbulent Stirring

The random velocity of dust aggregates increases owing to the gas drag from the turbulent velocity field as (Paper I)

Equation (34)

where $S={\rm{\Omega }}{t}_{{\rm{s}}}$ is the Stokes number, ${v}_{{\rm{t}}}=\sqrt{\alpha }{c}_{{\rm{s}}}$ is the magnitude of the turbulent velocity, α is the dimensionless turbulence strength (Cuzzi et al. 2001), and ${\tau }_{e}{{\rm{\Omega }}}^{-1}$ is the eddy turnover time. In the fiducial model, we adopt ${\tau }_{e}=1$ (Youdin 2011). For the isotropic turbulent velocity, the heating rate of ${\sigma }_{e}$ would be twice as large as that of ${\sigma }_{i}$, and thus we adopt the following formulae (Kobayashi et al. 2016):

Equation (35)

Equation (36)

3.3.3. Turbulent Scattering

The fluctuations in gas density due to turbulence result in gravitational scattering of the dust aggregates. Okuzumi & Ormel (2013) derived the stirring rate of ${\sigma }_{e}$ due to this effect:

Equation (37)

where Cturb is a nondimensional coefficient. From the semianalytical discussion, the nondimensional coefficient was obtained as (Okuzumi & Hirose 2011; Gressel et al. 2012; Okuzumi & Ormel 2013)

Equation (38)

where H is the gas scale height, ${H}_{\mathrm{res},0}$ is the vertical half-width of the dead zone for the magnetorotational instability, and ${ \mathcal L }$ is a nondimensional saturation limiter that is less than or equal to unity. For simplicity, we set ${ \mathcal L }=1$. In the fiducial model, we adopt ${H}_{\mathrm{res},0}=H$, which leads to ${C}_{\mathrm{turb}}=3.1\times {10}^{-2}$. In Section 5, we discuss the effect of Cturb. The stirring rate of ${\sigma }_{i}$ is

Equation (39)

where ${\epsilon }_{i}$ is a nondimensional coefficient and is smaller than unity (Yang et al. 2012). We adopt ${\epsilon }_{i}=0.1$ (Kobayashi et al. 2016).

3.4. Equilibrium Random Velocity

Considering the above processes, we obtain the evolution equations for ${\sigma }_{e}$ and ${\sigma }_{i}$:

Equation (40)

Equation (41)

If the relaxation of ${\sigma }_{e}$ and ${\sigma }_{i}$ is sufficiently fast, we can obtain the equilibrium values of ${\sigma }_{e}$ and ${\sigma }_{i}$ from $d{\sigma }_{e}^{2}/{dt}=0$ and $d{\sigma }_{i}^{2}/{dt}=0$. In Section 4, we discuss the GI using the equilibrium values of ${\sigma }_{e}$ and ${\sigma }_{i};$ however, we note that the validity of this treatment is not trivial. The nonequilibrium effect is discussed in Section 5.

3.5. Condition for GI

For a dust layer that is perfectly coupled with an incompressible fluid, the Roche density ${\rho }_{{\rm{R}}}$ is often used for the GI condition (Sekiya 1983; Yamoto & Sekiya 2004). In this case, a buckling mode develops, and dust-rich gas clumps form. Planetesimals may form in these dense clumps (Sekiya 1983; Cuzzi et al. 2008). However, in this paper, we focus on large dust aggregates that have a large Stokes number. In other words, dust aggregates can move relative to the gas, and dust can collapse through the gas, where the Roche criterion may not be applicable. In this case, Toomre's Q better describes the GI condition (Toomre 1964). We approximate the dust layer as a fluid with a finite velocity dispersion and calculate Q as

Equation (42)

where ${v}_{x}={({v}_{{\rm{K}}}^{2}{\sigma }_{e}^{2}/2)}^{1/2}$ is the radial component of the dust velocity dispersion and CT is a nondimensional constant. The values of CT are ${C}_{{\rm{T}}}=\pi $ for an ideal gas and ${C}_{{\rm{T}}}=3.36$ for collisionless particles (Toomre 1964); we adopted ${C}_{{\rm{T}}}=3.36$. In the majority of this paper, we adopt the condition for the GI as $Q\lt {Q}_{\mathrm{cr}}\simeq 2$ (Paper I).

For comparison, we also examine the Roche criterion. For the uniform dust layer perfectly coupled with the incompressible gas, the Roche density is (Sekiya 1983)

Equation (43)

We adopt this as the Roche density. Strictly speaking, the coefficient of the Roche density depends on the vertical structure and the equation of state of the dust layer. For instance, the coefficient for the Gaussian dust density distribution is 0.78 (Yamoto & Sekiya 2004). Furthermore, in this paper we do not consider the incompressible gas. Thus, Equation (43) is considered as an order-of-magnitude estimate of the Roche density for loosely coupling dust. From Equation (43) we define the nondimensional value QR (e.g., Youdin 2011),

Equation (44)

where ${h}_{{\rm{d}}}=a{\sigma }_{i}$ is the dust scale height and ${h}_{\mathrm{cr}}={{\rm{\Sigma }}}_{{\rm{d}}}/(\sqrt{\pi }{\rho }_{{\rm{R}}})$ is the critical scale height. If ${Q}_{{\rm{R}}}\lesssim 1$, the dust layer tends to be unstable in the sense of the Roche criterion.

4. Condition for GI of Dust Aggregates

From md and ${\rho }_{\mathrm{int}}$, we can calculate the equilibrium value of ${\sigma }_{e}$ and Q and draw the GI region on the md${\rho }_{\mathrm{int}}$ plane where $Q\lt {Q}_{\mathrm{cr}}$. We will show that a moderate mass is favorable for the GI. The existence and shape of the GI region depend on the various disk parameters. In this section, we examine a general condition for the existence of a GI region in the md${\rho }_{\mathrm{int}}$ plane, independent of the evolution of dust aggregates. Whether an actual dust disk becomes gravitationally unstable depends on the mass–radius relation of the dust aggregates. This issue will be discussed in Section 5, where we consider the evolution of the dust.

4.1. GI Region in the Mass–Density Plane

We begin by considering the equilibrium state. We numerically calculate the equilibrium values of ${\sigma }_{e}$ and ${\sigma }_{i}$ by setting the right-hand sides of Equations (40) and (41) equal to zero. Then, from Equation (42), we calculate Q.

Figure 2 shows the equilibrium ${\sigma }_{e}$ and ${\sigma }_{i}$ for the fiducial model with $a=5\,\mathrm{au}$, $\alpha ={10}^{-3}$, and ${f}_{{\rm{g}}}=1$. In this model, ${\sigma }_{e}$ and ${\sigma }_{i}$ range from ${10}^{-5}$ to ${10}^{-3}$. Basically ${\sigma }_{e}$ and ${\sigma }_{i}$ have similar dependencies on md and ${\rho }_{\mathrm{int}}$. In most of the parameter regimes, ${\sigma }_{i}$ is smaller than ${\sigma }_{e}$. The ratio of ${\sigma }_{i}$ to ${\sigma }_{e}$ is discussed in detail below. Around the region where ${m}_{{\rm{d}}}\sim {10}^{15}\,{\rm{g}}$ and ${\rho }_{\mathrm{int}}\sim {10}^{-4}\,{\rm{g}}\,{\mathrm{cm}}^{-3}$, ${\sigma }_{e}$ has the smallest value of about ${10}^{-5}$. The GI is likely to occur around this region in the fiducial model.

Figure 2.

Figure 2. Equilibrium values of (a) eccentricity and (b) inclination on the md${\rho }_{\mathrm{int}}$ plane for the fiducial model. The dashed, short-dashed, and dotted curves correspond to $3\times {10}^{-5}$, 10−4, and $3\times {10}^{-4}$, respectively.

Standard image High-resolution image

We investigate Q on the md${\rho }_{\mathrm{int}}$ plane. Figure 3 shows the result for the fiducial model. The minimum value of Q is at ${m}_{{\rm{d}}}=6.3\times {10}^{15}\,{\rm{g}}$ and ${\rho }_{\mathrm{int}}=1.4\times {10}^{-4}\,{\rm{g}}\,{\mathrm{cm}}^{-3}$, where ${Q}_{\min }\simeq 0.6$, which is sufficiently small to allow the GI. Note that $Q\lt {Q}_{\mathrm{cr}}$ when ${\rho }_{\mathrm{int}}=1\times {10}^{-7}$ to $1\,{\rm{g}}\,{\mathrm{cm}}^{-3}$. Therefore, when the density is in the realistic range, the GI condition is inevitably satisfied during the evolution of the dust.

Figure 3.

Figure 3. GI region on the md${\rho }_{\mathrm{int}}$ plane for the fiducial model. The solid, dashed, and short-dashed curves correspond to Q = 1, 2, and 4, respectively. The dotted line denotes the evolution track of the self-gravity compression. The dot-dashed curve corresponds to ${Q}_{{\rm{R}}}=2$.

Standard image High-resolution image

The main heating and cooling mechanisms of ${\sigma }_{e}$ and ${\sigma }_{i}$ are shown on the md${\rho }_{\mathrm{int}}$ plane for the fiducial model in Figure 4. In the region of low mass and low density, turbulent stirring is the dominant heating mechanism because the gas drag is strong. In the region of high mass and high density, turbulent scattering is the dominant heating mechanism, and in the region of high mass and low density, gravitational scattering is the dominant heating mechanism. The heating rate due to gravitational scattering is proportional to the scattering cross section, which is about ${({{Gm}}_{{\rm{d}}}/{v}_{\mathrm{ran}}^{2})}^{2}$ (e.g., Ida 1990). In the high-mass region, the random velocity is approximately given by the escape velocity because collisional damping and gravitational scattering are dominant. Thus, Figure 2 shows that in this region the random velocity decreases with decreasing ${\rho }_{\mathrm{int}}$. Therefore, gravitational scattering is stronger for the lower density if we fix the mass. The region of turbulent scattering for ${\sigma }_{i}$ is smaller than that for ${\sigma }_{e}$ because the scattering rate for ${\sigma }_{i}$ is smaller than it is for ${\sigma }_{e}$, as shown in Equation (39).

Figure 4.

Figure 4. Dominant heating and cooling mechanisms for ${\sigma }_{e}$ (left) and ${\sigma }_{i}$ (right) for the fiducial model. The filled and open symbols represent gas drag and collisional damping for the dominant cooling process, respectively. The squares, circles, and triangles represent gravitational scattering, turbulent stirring, and turbulent scattering, respectively. The shaded region denotes the GI region where $Q\lt {Q}_{\mathrm{cr}}$. The dashed and dotted lines represent the approximated instability condition described by Equations (46) and (47), respectively.

Standard image High-resolution image

We compare the Roche criterion with the Toomre criterion. In the fiducial model, there is no region for ${Q}_{{\rm{R}}}\lt 1$. As shown in Figure 3, there is the region for ${Q}_{{\rm{R}}}\lt 2$, but it is smaller than that for $Q\lt 2$. Thus, the Roche criterion is harder to satisfy than the Toomre criterion. The region for ${Q}_{{\rm{R}}}\lt 2$ relatively extends toward high ${\rho }_{\mathrm{int}}$. In this region, the main heating source is turbulent scattering. The heating rate of the inclination due to turbulent scattering is small, which leads to a thin dust layer. Thus, there the dust layer is more likely to be unstable in the sense of the Roche criterion. In the following we use the Toomre criterion, which is more optimistic.

4.2. Dynamical Properties of Dust Aggregates

We examine the dust aggregate dynamics in detail to clarify the physical background of the GI. Figure 5(a) shows the ratio of the random velocity to the surface escape velocity, ${v}_{\mathrm{esc}}=\sqrt{2{{Gm}}_{{\rm{d}}}/{r}_{{\rm{c}}}}$. If this ratio is less than unity, gravitational focusing is effective. In the low-mass region, the ratio is sufficiently larger than unity, which means that gravitational focusing is negligible. In this case, the collisional cross section is well approximated by the geometrical cross section, and the condition for runaway growth is not satisfied (e.g., Ohtsuki et al. 1993; Kokubo & Ida 1996). In the high-mass region, the escape velocity is comparable to the random velocity, and thus the collisional cross section is enhanced by a factor of about 2.

Figure 5.

Figure 5. Dynamical parameters of dust aggregates (a) ${v}_{\mathrm{ran}}/{v}_{\mathrm{esc}}$, (b) ${v}_{\mathrm{ran}}/\eta {v}_{{\rm{K}}}$, (c) ${v}_{\mathrm{ran}}/{v}_{{\rm{H}}}$, (d) ${\sigma }_{i}/{\sigma }_{e}$, (e) CD, and (f) S on the md${\rho }_{\mathrm{int}}$ plane for the fiducial model. The solid and dashed curves show $Q={Q}_{\mathrm{cr}}$ and unity, respectively. The dotted line in panel (e) shows $\mathrm{Re}\,=\,1$ (${C}_{{\rm{D}}}=27.6$).

Standard image High-resolution image

When calculating u (Equation (31)), if ${v}_{\mathrm{ran}}/(\eta {v}_{{\rm{K}}})$ is small, then vran is negligible. Figure 5(b) shows ${v}_{\mathrm{ran}}/(\eta {v}_{{\rm{K}}})$. Other than when the mass and density are both high, this ratio is less than unity, and we can safely adopt the approximation $u\simeq \eta {v}_{{\rm{K}}}$.

The ratio of the random velocity to the Hill velocity, ${v}_{{\rm{H}}}={r}_{{\rm{H}}}{\rm{\Omega }}$, determines the regime of gravitational scattering, that is, the shear-dominated or dispersion-dominated regime (Weidenschilling 1989). Figure 5(c) shows ${v}_{\mathrm{ran}}/{v}_{{\rm{H}}}$. In the entire region, vran is larger than vH, which indicates that the random velocity is dispersion dominated.

Figure 5(d) shows the ratio ${\sigma }_{i}/{\sigma }_{e}$, which is determined by the heating and cooling mechanisms. Figure 4 shows the main heating and cooling mechanisms on the md${\rho }_{\mathrm{int}}$ plane. For the region where the main heating and cooling mechanism is gravitational scattering and collisions, respectively, ${\sigma }_{i}/{\sigma }_{e}$ is about 0.4–0.6, because gravitational scattering results in ${\sigma }_{i}/{\sigma }_{e}\simeq 0.5$ (Ida & Makino 1992). In the region with turbulent drag and gas drag, ${\sigma }_{i}/{\sigma }_{e}$ is larger than unity, and the damping rate of ${\sigma }_{e}$ due to gas drag is larger than that of ${\sigma }_{i}$. Thus, ${\sigma }_{i}$ is larger than ${\sigma }_{e}$. The ratio ${\sigma }_{i}/{\sigma }_{e}$ is about 0.8 where turbulent drag and collisions are dominant. In the region where turbulent scattering and collisions prevail, ${\sigma }_{i}/{\sigma }_{e}$ is less than 0.2, and the heating rate of ${\sigma }_{i}$ due to turbulent scattering is smaller than that of ${\sigma }_{e}$. This leads to a smaller value for ${\sigma }_{i}/{\sigma }_{e}$ (Yang et al. 2012; Kobayashi et al. 2016).

Figure 5(e) shows the drag coefficient CD. There is no Epstein regime in Figure 5(e). The Epstein regime appears when we consider the small dust aggregates with ${m}_{{\rm{d}}}\lt {10}^{7}{\rm{g}}$. For $\mathrm{Re}\,\lt \,1$ (${C}_{{\rm{D}}}\gt 27.6$), the Stokes law is a good approximation. In the region where the mass is small and the density is large, the gas drag obeys the Stokes law. On the other hand, for massive dust aggregates, the gas drag obeys Newton's law. Thus, in this region, CD is almost constant, at about 0.4–0.5.

The Stokes number is shown in Figure 5(f). For the low-mass, low-density region, S is less than unity, which means that the dust is well coupled with the gas. In the GI region, S is sufficiently larger than unity that the coupling does not occur.

We now consider the main heating and cooling processes shown in Figure 4. In the low-mass, high-density region, gas drag is the main cooling mechanism, while in the high-mass, low-density region, collisions dominate. Gas drag obeys the Stokes law in the low-mass, high-density region, as shown in Figure 5(e). In this case, the damping rate due to gas drag is $\propto {C}_{{\rm{D}}}{r}_{{\rm{c}}}^{2}{/m}_{{\rm{d}}}\propto {r}_{{\rm{c}}}{/m}_{{\rm{d}}}\propto {m}_{{\rm{d}}}^{-2/3}{\rho }_{\mathrm{int}}^{-1/3}$. Similarly, neglecting gravitational focusing, the damping rate due to collisions is $\propto {r}_{{\rm{c}}}^{2}{/m}_{{\rm{d}}}\propto {m}_{{\rm{d}}}^{-1/3}{\rho }_{\mathrm{int}}^{-2/3}$. Thus, for larger md and smaller ${\rho }_{\mathrm{int}}$, damping by collisions is more important than that by gas drag.

The region where gas drag is dominant for ${\sigma }_{e}$ is larger than that for ${\sigma }_{i}$. As shown in Figure 5(b), in most areas, ${\sigma }_{e},{\sigma }_{i}\lt \eta $. In this case, the damping rate due to gas drag can be approximated as ${(d{\sigma }_{e}^{2}/{dt})}_{\mathrm{drag}}\simeq -(3/{t}_{{\rm{s}}}){\sigma }_{e}^{2}$ and ${(d{\sigma }_{i}^{2}/{dt})}_{\mathrm{drag}}\,\simeq -(1/{t}_{{\rm{s}}}){\sigma }_{i}^{2}$. The damping time for ${\sigma }_{e}$ is ${t}_{{\rm{s}}}/3$, which is shorter than that for ${\sigma }_{i}$. On the other hand, the damping times due to collisions for ${\sigma }_{e}$ and ${\sigma }_{i}$ are the same. Thus, the gas drag region for ${\sigma }_{e}$ is larger than that for ${\sigma }_{i}$.

4.3. Critical Turbulent Strength for GI

By a similar way to that used in Paper I, we derive the critical α for the existence of the GI region. First, we focus on the lower mass boundary of the GI region around ${\rho }_{\mathrm{int}}\simeq {10}^{-6}\,{\rm{g}}\,{\mathrm{cm}}^{-3}$. In the low-density region, the main heating and cooling mechanisms are turbulent stirring and collisions, respectively. In this case, the equilibrium ${\sigma }_{e}$ is approximated by ${(d{\sigma }_{e}^{2}/{dt})}_{\mathrm{turb},\mathrm{stir}}+{(d{\sigma }_{e}^{2}/{dt})}_{\mathrm{col}}\simeq 0$. We neglect the second term in Equation (25) since ${v}_{\mathrm{ran}}\gt {v}_{\mathrm{esc}}$ and gravitational focusing is not effective, as shown in Figure 5(a). Furthermore, as shown in Figure 5(d), ${\sigma }_{i}/{\sigma }_{e}$ is almost constant in the low-density region. Thus, assuming ${\sigma }_{i}/{\sigma }_{e}=0.71$, we evaluate the first term of Equation (25) and obtain ${P}_{\mathrm{col}}\simeq 2.13{\tilde{r}}_{{\rm{c}}}^{2}$. Substituting this into Equation (21), we obtain the approximated ${(d{\sigma }_{e}^{2}/{dt})}_{\mathrm{col}}$. Figure 5(f) shows $S\gg 1$, where we obtain ${(d{\sigma }_{e}^{2}/{dt})}_{\mathrm{turb},\mathrm{stir}}\simeq 4{\tau }_{e}{v}_{{\rm{t}}}^{2}{\rm{\Omega }}/3{v}_{{\rm{K}}}^{2}{S}^{2}$ from Equation (35). We find that ${v}_{\mathrm{ran}}\lt \eta {v}_{{\rm{K}}}$ in Figure 5(b). Thus, from Equation (31), we approximate $u\simeq \eta {v}_{{\rm{K}}}$. Substituting this into Equation (32), we calculate S and finally obtain the approximated ${(d{\sigma }_{e}^{2}/{dt})}_{\mathrm{turb},\mathrm{stir}}.$ From Figure 5(e), Newton's drag is a good approximation; thus, we assume ${C}_{{\rm{D}}}\simeq 0.5$. Based on these approximations, we calculate ${\sigma }_{e}$ and obtain

Equation (45)

From $Q\lt {Q}_{\mathrm{cr}}$, we obtain the inequality

Equation (46)

Next, we focus on the upper mass boundary of the GI region around ${\rho }_{\mathrm{int}}\simeq {10}^{-1}\,{\rm{g}}\,{\mathrm{cm}}^{-3}$. In the high-mass, high-density region, the main heating and cooling mechanisms are turbulent scattering and collisions, respectively. We obtain the equilibrium value for ${\sigma }_{e}$ from ${(d{\sigma }_{e}^{2}/{dt})}_{\mathrm{turb},\mathrm{grav}}+{(d{\sigma }_{e}^{2}/{dt})}_{\mathrm{col}}=0$. In evaluating ${(d{\sigma }_{e}^{2}/{dt})}_{\mathrm{col}}$, we adopt the same approximations described above. From $Q\lt {Q}_{\mathrm{cr}}$, we obtain the upper limit of md:

Equation (47)

In Figure 4, mlow and mhigh are plotted, and they are in rough agreement with the numerical results.

For the dust aggregates to trigger the GI, it is necessary that ${m}_{\mathrm{low}}\lt {m}_{\mathrm{high}}$. If this inequality is not satisfied, the GI region does not exist. Thus, we derive the following condition for the GI:

Equation (48)

Using the disk model, we rewrite ${\alpha }_{\mathrm{cr},1}$ as

Equation (49)

As shown in Section 5.1, this condition agrees well with the numerical results.

In the fiducial model (${\beta }_{{\rm{g}}}=3/2$ and ${\beta }_{{\rm{t}}}=3/7$), we find a weak dependence on a: $\propto {a}^{-1/14}$. Note that for the optically thin MMSN model (${\beta }_{{\rm{t}}}=1/2$ and ${\beta }_{{\rm{g}}}=3/2$) (Hayashi 1981; Hayashi et al. 1985), this dependence vanishes completely. Thus, for realistic disk models, the dependence on a is generally weak.

5. GI with Evolution of Dust

In the previous section, we investigated a condition on the dust aggregate that would bring about the GI; in this section, we examine whether the GI occurs as dust aggregates evolve in a protoplanetary disk.

5.1. Dependence on Disk Parameters

First, assuming the equilibrium random velocity, we examine the dependence of the critical α for the GI on the disk parameters. We adopt the dust evolution described in Section 2.2.2 and check whether its track crosses the GI region. Figure 6(a) shows the dependence on fg. The GI is more likely with larger fg and smaller α. The preference for small α occurs because turbulence is the main heating mechanism. From Equation (37), when the gas surface density (fg) is large, the heating rate due to turbulent scattering is also large. In our model, the dust surface density also increases with fg, since the dust-to-gas density ratio γ is fixed. When the dust surface density is large, this leads to strong self-gravity and a high collision frequency. Among these competing effects, the increases in self-gravity and collision frequency are dominant. Thus, the GI is more likely to occur when fg is large.

Figure 6.

Figure 6. Critical viscous parameter α for the GI vs. (a) fg, (b) a, (c) γ, (d) T1, (e) ${\tau }_{e}$, and (f) Cturb. All points show the existence of the GI region in the area ${10}^{8}\lt {m}_{{\rm{d}}}\lt {10}^{20}\,{\rm{g}}$ and ${10}^{-6}\,{\rm{g}}\,{\mathrm{cm}}^{-3}\lt {\rho }_{\mathrm{int}}\lt 1\,{\rm{g}}\,{\mathrm{cm}}^{-3}$. Open triangles indicate that the dust evolution does not cross the GI region. Open circles indicate where the dust evolution crosses the GI region with the equilibrium random velocity but not with the nonequilibrium random velocity, while filled circles show where the dust evolution crosses the GI region with both equilibrium and nonequilibrium random velocities. The solid and dashed curves show ${\alpha }_{\mathrm{cr},1}$ (Equation (49)) and ${\alpha }_{\mathrm{cr},2}$ (Equation (56)), respectively.

Standard image High-resolution image

We found that the condition for the existence of the GI region (Equation (49)) agrees well with the numerical results for ${f}_{{\rm{g}}}\gt 0.3$, while it overestimates α for ${f}_{{\rm{g}}}\lt 0.3$. To determine the reason for this, in Figure 7 we plotted the main heating and cooling mechanisms for $\alpha ={10}^{-4}$ and ${f}_{{\rm{g}}}=0.1$. For low fg, turbulent scattering is insignificant because its heating rate is proportional to ${{\rm{\Sigma }}}_{{\rm{g}}}^{2}$. In deriving Equation (49), we assumed that the upper boundary of the GI region is determined by a balance between turbulent scattering and collisions. However, this assumption breaks down, and thus, in this parameter region, Equation (49) differs from the numerical results.

Figure 7.

Figure 7. Main heating and cooling mechanisms for ${\sigma }_{e}$ with $\alpha =1\times {10}^{-4}$ and ${f}_{{\rm{g}}}=0.1$. The other parameters are the same as those of the fiducial model. The symbols are the same as those in Figure 4. The shaded region is the GI region.

Standard image High-resolution image

As expected from Equation (49), the dependence on a is weak. From the numerical results shown in Figure 6(b), the existence of the GI region is independent of a. The value of α required to cross the GI region slightly decreases with a, but its dependence is very weak.

As shown in Figure 6(c), the dependence on the dust-to-gas ratio γ is strong. The GI easily takes place for larger values of γ. The critical α for the existence of the GI region, ${\alpha }_{\mathrm{cr},1}$ (Equation (49)), is proportional to ${\gamma }^{3}$, which agrees well with the numerical results. However, the dependence of the critical α on γ for crossing the GI region is different; it is approximately proportional to ${\gamma }^{2}$. This dependence will be discussed below. The difference from ${\alpha }_{\mathrm{cr},1}$ increases with γ for $\gamma \gtrsim 0.05$.

The critical α decreases with increasing T1, as shown in Figure 6(d). Increased temperatures lead to suppression of the GI. From Equation (4), η is proportional to T1. The typical difference between the velocity of dust and that of gas is determined from $\eta {v}_{{\rm{K}}}$ because ${v}_{\mathrm{ran}}\lt \eta {v}_{{\rm{K}}}$. Thus, a higher T1 means that the gas drag is strong, and this causes strong turbulent stirring and inhibits the GI. When ${T}_{1}=280$, the critical α that is often adopted is 0.4 times that when ${T}_{1}=120$.

Figure 6(e) shows that the critical α decreases with increasing ${\tau }_{e}$ as ${\tau }_{e}^{-1/2}$. A small ${\tau }_{e}$ means that the duration of the fluctuations in the turbulent velocity is short. In this case, from Equation (35), the heating rate due to turbulent stirring is small. Thus, the GI tends to occur for smaller ${\tau }_{e}$.

The critical α decreases with increasing Cturb as ${C}_{\mathrm{turb}}^{-1/2}$, as shown in Figure 6(f). From Equation (37), the heating rate due to turbulent scattering decreases with increasing Cturb. Thus, for smaller Cturb, the critical α is larger.

Figure 8 shows the influence of the power-law indices ${\beta }_{{\rm{g}}}$ and ${\beta }_{{\rm{t}}}$. In the fiducial model, the dependence on a is weak. However, if we adopt the general power-law indices, the dependence on a can be strong. As discussed in Section 4.3, the dependence on a vanishes only if ${\beta }_{{\rm{g}}}-{\beta }_{{\rm{t}}}=1$. As ${\beta }_{{\rm{g}}}$ increases, the dependence on a becomes weak, while as ${\beta }_{{\rm{t}}}$ increases, the dependence on a becomes strong. This behavior is consistent with Equation (49).

Figure 8.

Figure 8. Critical viscous parameter α for the GI vs. a for (a) ${\beta }_{{\rm{g}}}=0.5$, (b) ${\beta }_{{\rm{t}}}=0.5$, (c) ${\beta }_{{\rm{g}}}=1.0$, (d) ${\beta }_{{\rm{t}}}=1.0$, (e) ${\beta }_{{\rm{g}}}=1.5$, and (f) ${\beta }_{{\rm{t}}}=1.5$. The other parameters are the same as those in the fiducial model. The symbols are the same as those in Figure 6.

Standard image High-resolution image

5.2. Effect of Nonequilibrium Random Velocity

In this section, we clarify the effect of the nonequilibrium random velocity on the onset of the GI. In the previous section, we assumed the equilibrium ${\sigma }_{e}$ and ${\sigma }_{i}$. However, under some conditions, the relaxation time of ${\sigma }_{e}$ and ${\sigma }_{i}$ may be comparable to the growth time of dust aggregates. In this case, the equilibrium values change before ${\sigma }_{e}$ and ${\sigma }_{i}$ reach equilibrium. Thus, the actual ${\sigma }_{e}$ and ${\sigma }_{i}$ may deviate from the equilibrium values.

We will simultaneously consider the evolution of the mass and random velocity. The mass evolution equation is

Equation (50)

where Cgrow is the sticking probability. The perfect accretion corresponds to ${C}_{\mathrm{grow}}=1$. For ${C}_{\mathrm{grow}}\ne 1$, we may need to change Ccol. Strictly speaking, Ccol depends on the energy dissipation rate on collisions, such as the restitution coefficient and the friction coefficient on the dust aggregate surface. For simplicity we assume ${C}_{\mathrm{col}}=1/2$ for any Cgrow. We solve the differential equations given as Equations (40), (41), and (50) for ${\sigma }_{e}$, ${\sigma }_{i}$, and md, respectively. For simplicity, we assume that dust aggregates always have the equilibrium internal density, ${\rho }_{\mathrm{int}}={\rho }_{\mathrm{eq}}$.

Figure 9 shows the evolution of Q for the fiducial model with an initial mass of $m={10}^{10}\,{\rm{g}}$ and initial equilibrium values for ${\sigma }_{e}$ and ${\sigma }_{i}$. We assume the perfect accretion (${C}_{\mathrm{grow}}=1$). As shown below, in the case of the perfect accretion, the difference between the equilibrium and nonequilibrium models is the most significant. Toomre's Q deviates slightly from the equilibrium value since the dust aggregates grow during the relaxation of the random velocity. Among the models shown in Figure 6, the deviation of the minimum Q is 9.7% ± 13.2%. Thus, the nonequilibrium effect is not significant for the onset of the GI. We also evaluated the influence of the initial values by considering values for ${\sigma }_{e}$ and ${\sigma }_{i}$ that were 5 times and 1/5 times their equilibrium values. As shown in Figure 9, the difference in Q that stems from the initial values quickly vanishes as the system evolves.

Figure 9.

Figure 9. Evolution of Q as the dust evolves in the fiducial model with the nonequilibrium and equilibrium (solid curve) random velocity. For the nonequilibrium model, the initial ${\sigma }_{e}$ and ${\sigma }_{i}$ are 1/5 (short-dashed), 1 (long-dashed), and 5 (dotted) times as large as the equilibrium values.

Standard image High-resolution image

We compare the various relevant timescales. The growth timescale is defined as

Equation (51)

and the timescale of gravitational scattering is defined as

Equation (52)

The other dynamical timescales tcol, ${t}_{\mathrm{gas},\mathrm{drag}}$, ${t}_{\mathrm{turb},\mathrm{stir}}$, and ${t}_{\mathrm{turb},\mathrm{grav}}$ are defined in the same way. The left panel of Figure 10 shows these timescales. If the growth is sufficiently slow compared with the dynamical timescales, the eccentricity reaches the equilibrium value. Figure 10 shows that the growth timescale is comparable to the dynamical timescales. Therefore the equilibrium eccentricity, which depends on the mass, changes before the eccentricity converges to the equilibrium value. However, since the growth timescale is not very different from the dynamical timescales, the gap from the equilibrium value is not so large.

Figure 10.

Figure 10. Various timescales (left) and the ratio of the inclination to the eccentricity (right) against the mass. The timescales are tgrow (dot-dashed), tcol (thick solid), ${t}_{\mathrm{gas},\mathrm{drag}}$ (thick dashed), tgrav (thin solid), ${t}_{\mathrm{turb},\mathrm{stir}}$ (thin dashed), and ${t}_{\mathrm{turb},\mathrm{grav}}$ (thin dotted), respectively. The initial ${\sigma }_{e}$ and ${\sigma }_{i}$ are set as the equilibrium values.

Standard image High-resolution image

Figure 10 shows two sharp peaks of tgrav. This is because PVS is negative if ${\sigma }_{i}/{\sigma }_{e}$ is small (Ohtsuki et al. 2002). In the relatively high velocity cases, gravitational scattering tends to realize ${\sigma }_{i}/{\sigma }_{e}\simeq 0.5$ (Ida 1990). Therefore, low ${\sigma }_{i}/{\sigma }_{e}$ leads to negative PVS, while high ${\sigma }_{i}/{\sigma }_{e}$ leads to negative QVS. The right panel of Figure 10 shows that if ${\sigma }_{i}/{\sigma }_{e}$ is about 0.3 around ${m}_{{\rm{d}}}\sim {10}^{16}\,{\rm{g}}$, then PVS is negative. Around the points where ${P}_{\mathrm{VS}}\simeq 0$, tgrav becomes very large.

The effect of the imperfect accretion is shown in Figure 11. The initial ${\sigma }_{e}$ and ${\sigma }_{i}$ are set as the equilibrium values. The evolution timescale is the fastest for the perfect accretion model. As Cgrow decreases, the evolution timescale becomes longer, which is inversely proportional to Cgrow. The difference between the equilibrium and nonequilibrium models becomes less with decreasing Cgrow. This is because for smaller Cgrow, the growth time is longer compared to the dynamical timescales, and thus the eccentricity can converge to the equilibrium value before the mass changes.

Figure 11.

Figure 11. Time evolution of Q for ${C}_{\mathrm{grow}}=0.1$ (red), 0.3 (blue), and 1.0 (black) in the fiducial model. The solid curves correspond to the nonequilibrium models, and the dotted curves correspond to the equilibrium models.

Standard image High-resolution image

In Figures 6 and 8, we examine the effect of the nonequilibrium velocity on the GI. If the nonequilibrium effect is considered, the value of α necessary for crossing the GI region decreases slightly; however, the difference is small. The nonequilibrium effect of the random velocity is thus insignificant. In summary, the above results justify the equilibrium random velocity model.

5.3. Condition for Crossing the GI Region

The dust evolution track is characterized by the monomer parameters Eroll, r0, and ${\rho }_{0}$. We first examine the effect of Eroll. In Figure 12, it can be seen that Eroll has little effect on the critical α for crossing the GI region. For small Eroll, the critical α only slightly decreases. This is due to the structure of the GI region. As shown in Figure 13, the GI region stretches as α decreases. In particular, the upper bound on the GI region increases rapidly with α. For $\alpha =2\times {10}^{-3}$, we can see the elongated GI region for ${\rho }_{\mathrm{int}}\gt 0.1\,{\rm{g}}\,{\mathrm{cm}}^{-3}$. If this region appears, the dust evolution inevitably crosses the GI region.

Figure 12.

Figure 12. Same as Figure 6, but vs. Eroll.

Standard image High-resolution image
Figure 13.

Figure 13. GI region for $\alpha =2\times {10}^{-3}$ (solid), $4\times {10}^{-3}$ (long-dashed), and $6\times {10}^{-3}$ (short-dashed). The dotted lines show the dust evolution with ${E}_{\mathrm{roll}}=4.7\times {10}^{-12}\,\mathrm{erg},4.7\times {10}^{-9}\,\mathrm{erg},$ and $4.7\times {10}^{-6}\,\mathrm{erg}$ from top to bottom, respectively.

Standard image High-resolution image

To examine the condition for the existence of an elongated GI region, in Figure 14 we show the main heating and cooling mechanisms for $\alpha =2\times {10}^{-3}$. The lower boundary of the elongated region is determined by the balance between turbulent stirring and gas drag. We calculate ${(d{\sigma }_{e}^{2}/{dt})}_{\mathrm{turb},\mathrm{stir}}$ in the same way as in Section 4.3 except for CD. Figure 5(e) shows that Stokes drag ${C}_{{\rm{D}}}\simeq 24/\mathrm{Re}$ is a good approximation. Thus, we adopt Stokes drag. Using these approximations, we calculate Q. From $Q\lt 2$, we obtain

Equation (53)

Similarly, the upper boundary of this region is determined by the balance between turbulent scattering and gas drag. Thus, we obtain the upper boundary as

Equation (54)

The condition for the existence of an elongated region is ${m}_{\mathrm{low},2}\lt {m}_{\mathrm{high},2}$, from which we obtain the critical α for crossing the GI region:

Equation (55)

which can be rewritten as

Equation (56)

We show this condition in Figure 6. If $\alpha \lt {\alpha }_{\mathrm{cr},2}$, the dust evolution crosses the GI region. Thus, we propose the following condition for the onset of the GI:

Equation (57)

Figure 14.

Figure 14. Main heating and cooling mechanisms for ${\sigma }_{e}$ with $\alpha =2\times {10}^{-3}$. The other parameters are the same as those of the fiducial model. The symbols are the same as in Figure 4. The dashed and dotted lines indicate the approximated boundaries of the elongated GI region described by Equations (53) and (54), respectively.

Standard image High-resolution image

In deriving this condition, we did not consider the nonequilibrium effect of the random velocity. However, as shown in Figure 6, that effect is insignificant. For the outer disk ($a=10\mbox{--}20\,\mathrm{au}$), we need a safety factor, such as $\alpha \lt 0.5\min ({\alpha }_{\mathrm{cr},1},{\alpha }_{\mathrm{cr},2})$.

In Paper I, we proposed a similar condition, and its parameter dependence is exactly the same as that of ${\alpha }_{\mathrm{cr},1}$. Other than when γ is particularly large or particularly small, the difference between ${\alpha }_{\mathrm{cr},1}$ and ${\alpha }_{\mathrm{cr},2}$ is very small. Thus, the simple condition proposed in Paper I ($0.5\times {\alpha }_{\mathrm{cr},1}$) is also a good estimate of the condition for crossing the GI region in most parameter regimes.

6. Summary and Discussion

We have investigated the stability of a disk that consists of porous icy dust aggregates in a turbulent gas disk. We calculated the random velocity of the aggregates, taking into account gravitational scattering, collisions, gas drag, turbulent stirring, and turbulent scattering. In Paper I, we assumed an isotropic velocity dispersion and an equilibrium random velocity of the aggregates. In this paper, we removed these assumptions. We separately calculated the evolution of the eccentricity and that of the inclination. We found that under the evolution of dust by self-gravity compression, the GI is inevitable when the disk parameters are in a realistic range. In the MMSN model, the GI takes place when the turbulent viscosity parameter α is less than about $4\times {10}^{-3}$. We estimated the critical α and confirmed that it is in good agreement with numerical results. Thus, this estimate can be applied to general disk models.

In this paper, we adopted the equal-mass dust aggregates for simplicity. This assumption breaks down in some parameter ranges, and a size distribution develops. A wide or bimodal size distribution of aggregates may alter the heating and cooling rates quantitatively. Furthermore, we need to include an additional dynamical effect, dynamical friction between different-sized aggregates. These effects are beyond the scope the present paper and should be investigated separately.

Finally, we consider the post-GI evolution. From linear analyses of the axisymmetric mode of the GI, the instability condition is $Q\lt 1$ (Toomre 1964; Goldreich & Lynden-Bell 1965a). However, as pointed out by Michikoshi et al. (2007, 2009, 2010), the axisymmetric mode does not appear in the dust disk. This is because the nonaxisymmetric mode (gravitational wake) grows for $1\lesssim Q\lesssim 2$, due to the swing amplification mechanism (Goldreich & Lynden-Bell 1965b; Julian & Toomre 1966; Toomre 1981; Fuchs 2001; Michikoshi & Kokubo 2016a, 2016b). Initially, Q of the porous dust disk is sufficiently larger than 2, and thus the disk is stable in any mode. As the dust aggregate evolves as a result of self-gravity compression, Q decreases gradually. For sufficiently small α, Q finally becomes less than 2, and gravitational wakes appear. If the energy dissipation is effective and Q decreases rapidly compared to the dynamical timescale, Q may become less than unity, which leads to the development of the axisymmetric mode. However, such a rapid decrease of Q is unlikely. The formation of a gravitational wake has also been confirmed by the hydrodynamics simulation of a dust layer (Wakita & Sekiya 2008). Michikoshi et al. (2007) conducted N-body simulations that showed that the gravitational wakes fragment to form planetesimals. However, this fragmentation does not always take place. For example, in a system where gas drag and inelastic collisions among particles are not effective, such as in a collisionless system, gravitational wakes develop as a result of the GI, but they do not fragment to form gravitationally bound objects (Toomre & Kalnajs 1991; Fuchs et al. 2005; Michikoshi & Kokubo 2014, 2016a). Therefore, energy dissipation in the wake is essential for planetesimal formation, and this suggests the existence of an additional condition for planetesimal formation. The energy dissipation rate may be a key parameter, as it is in a gas disk, where if the cooling timescale is comparable to or shorter than the dynamical timescale, the disk fragments (Gammie 2001). In our next paper, we will examine the post-GI evolution and scrutinize the necessary conditions for the formation of planetesimals.

The authors would like to thank the anonymous referee for useful comments.

Please wait… references are loading.
10.3847/1538-4357/aa7388