Rotational Disruption of Dust and Ice by Radiative Torques in Protoplanetary Disks and the Implications for Observations

and

Published 2020 September 16 © 2020. The American Astronomical Society. All rights reserved.
, , Citation Ngo-Duy Tung and Thiem Hoang 2020 ApJ 901 6 DOI 10.3847/1538-4357/abacbf

Download Article PDF
DownloadArticle ePub

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

0004-637X/901/1/6

Abstract

Dust and ice mantles on dust grains play an important role in various processes in protoplanetary disks (PPDs) around a young star, including planetesimal formation, surface chemistry, and being the reservoir of water in habitable zones. In this paper, we perform two-dimensional modeling of rotational disruption of dust grains and ice mantles due to centrifugal force within suprathermally rotating grains spun-up by radiative torques for disks around T-Tauri and Herbig Ae/Be stars. We first study rotational disruption of large composite grains and find that large aggregates could be disrupted into individual nanoparticles via the RAdiative Torque Disruption (RATD) mechanism. We then study rotational desorption of ice mantles and ro-thermal desorption of molecules from the ice mantle. We find that ice mantles in the disk's warm surface layer and above can be disrupted into small icy fragments, followed by rapid evaporation of molecules. We suggest that the rotational disruption mechanism can replenish the ubiquitous presence of polycyclic aromatic hydrocarbons/nanoparticles in the hot surface layers of PPDs as observed in mid-IR emission, which are presumably destroyed by extreme ultraviolet (UV) stellar photons. We find that the water snowline is more extended in the presence of rotational desorption, which would decrease the number of comets but increase the number of asteroids formed in the solar nebula. Finally, we suggest that RATD breaking up carbonaceous grains more efficiently than it does silicates might resolve the carbon deficit problem measured on the Earth and rocky bodies.

Export citation and abstract BibTeX RIS

1. Introduction

Dust and ice mantles on dust grains play an important role in various processes in protoplanetary disks (PPDs) around young stars, including the formation of planetesimals (Blum & Wurm 2008), surface chemistry (Henning & Semenov 2013), and being the reservoir of water in habitable zones. Ice mantles also affect the formation of super-Earth (Howe & Burrows 2015) and giant-planet cores (Kennedy & Kenyon 2008). In particular, the desorption of water and volatiles from ice mantles affects the chemical composition of giant planet atmospheres (Öberg et al. 2011; Madhusudhan et al. 2014) and the delivery of water to the surfaces of terrestrial planets (Raymond et al. 2007).

PPDs exhibit a strong gradient of local density and temperatures in the vertical and radial directions, due to the effect of stellar radiation and gravity. Near the star, water ice sublimates within a region of temperatures of ${T}_{\mathrm{sub}}\gtrsim 150\,{\rm{K}}$, and the boundary defines the water snowline about 2.7 au from a Sun-like star (Hayashi 1981). The physical structure of a PPD can be characterized by three distinct regions, including the hot upper layer directly irradiated by stellar and interstellar photons, partly shielded warm intermediate layer, and the disk interior near the midplane, which is completely shielded from stellar photons (see, e.g., Henning & Semenov 2013).

Very large molecules, such as polycyclic aromatic hydrocarbon (PAH) molecules, are usually detected from the surface layers of disks around Herbig Ae/Be stars and some T-Tauri stars (Habart et al. 2004; Seok & Li 2017) via mid-infrared emission features at 3.3, 6.2, 7.7, 8.6, 11.3, and 17 μm (Leger & Puget 1984; Allamandola et al. 1985; Draine & Li 2007; Smith et al. 2007). Furthermore, near-infrared (near-IR) observations from the Very Large Telescope (VLT) reveal the presence of carbonaceous nanoparticles throughout the surface of PPDs, even in the central cavity where large grains are depleted due to grain growth (Boutéraon et al. 2019). Moreover, microwave observations reveal the existence of nanoparticles, including PAHs, nanosilicates, and nanodiamonds, which are explained by spinning dust emission (Hoang et al. 2018).

The widespread presence of PAHs/nanoparticles in the surface layer is unexpected because those nanoparticles are thought to be efficiently destroyed by extreme UV and X-ray radiation from the central star (Siebenmorgen & Krügel 2010). Thus, the origin of such PAHs/nanoparticles in the disk surface layer remains unclear. Previously, PAHs have been suggested to form in the regions of high temperature and density behind the rim of PPDs (see Kamp 2011). An alternative explanation is that nanoparticles may follow an evolution different from that of classical grains (size of 0.1 μm). As a result, while the classical grains are depleted in the disk, due to coagulation and settling, PAHs/nanoparticles that are well-mixed into the gas can be exempt from grain settling and coagulation, and turbulence mixing can frequently transport nanoparticles from the disk interior to the surface (see Dullemond et al. 2007). In this paper, we will show that PAHs/nanoparticles can be reproduced by rotational disruption of dust grains (a ≳ 0.1 μm) via the RAdiative Torque Disruption (RATD) mechanism, as discovered by Hoang et al. (2019).

Destruction of ice mantles essentially occurs in the hot surface layer via thermal sublimation where ice mantles are heated to above $150\,{\rm{K}}$. In the warm molecular layer, with grain temperatures of ${T}_{d}\sim 30\mbox{--}100\,{\rm{K}}$, ice mantles can survive against thermal sublimation and are usually thought to play a central role for formation and desorption of complex molecules (see, e.g., Henning & Semenov 2013). Indeed, ice mantles are considered the main route to form complex organic molecules (COMs), i.e., organic molecules that contain more than six atoms, such as CH3OH, HCOOH, CH3CHO, and C2H5OH; see Herbst & van Dishoeck (2009) for a review. To date, several complex molecules, including CH3CN (Öberg et al. 2015), CH3OH (Walsh et al. 2016), and formic acid (HCOOH; Favre et al. 2018) have been detected from PPDs. Interestingly, these molecules are observed at distant locations of R > 10 au where grain temperatures are lower than the sublimation threshold of observed molecules.4 The question is: how can such complex molecules be released to the gas at large distances?

Recently, Hoang & Tram (2020) showed that the entire icy grain mantle could be disrupted due to centrifugal stress induced by suprathermal rotation of grains spun-up by radiative torques (RATs); see Hoang (2020) for a review. Depending on the local gas density, rotational desorption of ice can occur at temperatures much lower than the water sublimation limit at $\sim 150\,{\rm{K}}$. Subsequently, water and complex molecules rapidly desorb from the resulting icy fragments due to thermal spikes. This rotational disruption mechanism is found to release water and complex molecules at much lower temperatures than previously predicted by thermal sublimation. Additionally, Hoang & Tung (2019) discovered that centrifugal potential induced by grain suprathermal rotation acts to reduce the potential barrier of adsorbed molecules onto the ice mantle and significantly enhances the rate of desorption. This mechanism is termed ro-thermal desorption.5 As a result, we expect that these mechanisms would be efficient in the disk conditions even in the warm layer, which would dramatically affect the physical and chemical properties of PPDs. In this paper, we will apply these mechanisms to perform two-dimensional modeling of ice evolution and molecule desorption for PPDs around T-Tauri and Herbig Ae/Be stars. Moreover, PAHs/nanoparticles locked up in the ice mantle could also be released by rotational and ro-thermal desorption of ice mantles (Hoang & Tung 2019).

The structure of this paper is as follows. In Section 2, we will describe a physical model of PPDs. In Section 3, we will briefly review the physical mechanisms of RATD, rotational desorption of ice mantles, and ro-thermal desorption, respectively. We will then apply these mechanisms for the PPDs' conditions and calculate disruption and desorption sizes of grains in Section 4. We study the effect of rotational disruption on absorption and scattering opacity in Section 5. Section 6 is devoted to discussing the implications of the applied mechanism on the detection of COMs and carbonaceous nanoparticles in the disks. We give a summary of our main findings in Section 7.

2. Physical Model of a PPD

2.1. A Passive Irradiated Disk Model

For our calculations, we adopt a two-dimensional (2D) flared, radiative, and hydrostatic equilibrium disk model (Chiang & Goldreich 1997) as illustrated in Figure 1. The disk surface is defined by a slant path of unity optical depth of τV ∼ 1. This layer is directly heated by stellar radiation, and is considered a photodissociated region. Dust and gas below this layer are heated by attenuated stellar radiation and define a warm layer ($T\sim 30\mbox{--}100\,{\rm{K}})$. The final region is the disk interior, which is completely shielded from stellar radiation and only heated by infrared emission from hot dust in the surface layer.

Figure 1.

Figure 1. Schematic illustration of a flared disk around a young star, consisting of three vertical layers: hot surface layer, warm intermediate layer, and disk interior.

Standard image High-resolution image

2.2. Gas Density Profile

The gas number density at disk radius R and height z for the hydrostatic disk model, assuming a Gaussian vertical profile (Lynden-Bell & Pringle 1974), is given by

Equation (1)

where Σ(R) is the total surface mass density at radius R given by

Equation (2)

where α is the model constant, and Σ1 is the surface mass density at $R=1\,\mathrm{au}$.

The pressure height scale Hp is described by

Equation (3)

where H0 is the aspect ratio at the reference radius R0.

For R0 = 100 au, H0/R0 is taken to be 0.1 as a fiducial model, which corresponds to Hp/R = 0.1 × 31/7 at Rout = 300 au. The chosen aspect ratio is much lower than predicted by Chiang & Goldreich (1997), but is comparable to observations (Avenhaus et al. 2018). Here, we assume α = 1 and Σ1 is varied to cover a wide range of disk mass. Other physical parameters are listed in Table 1, including the stellar temperature T, stellar mass M, stellar radius R, and the inner and outer disk radii Rin and Rout.

Table 1.  Physical Parameters of a PPD

Objects T M R Rin Rout
  $({\rm{K}})$ (M) (R) $(\mathrm{au})$ $(\mathrm{au})$
Herbig Ae/Be 10000 2 2 1 300
T-Tauri 4000 0.5 2 0.1 300

Download table as:  ASCIITypeset image

2.3. Grain Temperature and Radiation Field

With the disk model defined, we compute the grain temperature throughout the disk by performing a radiative transfer calculation using the publicly available 3D Monte Carlo radiative transfer code (radmc-3d; Dullemond et al. 2012).6 The grid resolutions are Nr = 128 for Rin < R < Rout and Nθ = 128 for 0 < θ < 2π. The number of photons is Nphot ∼ 109. The dust opacity is calculated assuming a power-law size distribution, n(a) ∝ aq (amin < a < amax) with the absorption cross section for spherical grains computed using the Mie theory coded from Bohren & Huffman (1983), assuming the optical constant of amorphous silicate (Mg0.7Fe0.3SiO3).7 We adopt lower and upper cutoffs for the size distribution of large grains amin = 0.1 μm and amax = 10 μm, and the standard MRN distribution q = 3.5 (Mathis et al. 1983). The dust has a total mass ranging from 2 × 10−6 to 2 × 10−3M for ${{\rm{\Sigma }}}_{1}=1\mbox{--}1000\,{\rm{g}}\,{\mathrm{cm}}^{-3}$ and the same vertical structure as the gas.

Figures 2 and 3 show the gas density and grain temperature from radmc-3d for the disks around T-Tauri and Herbig Ae/Be stars, respectively, with ${{\rm{\Sigma }}}_{1}=1000\,{\rm{g}}\,{\mathrm{cm}}^{-2}$.

Figure 2.

Figure 2. Gas density (left panel) and grain temperature obtained from radmc-3d (right panel) for a PPD around T-Tauri stars as a function of the disk radius R and disk height scaled by radius z/R with ${{\rm{\Sigma }}}_{1}=1000\,{\rm{g}}\,{\mathrm{cm}}^{-2}$. Black lines show the grain temperatures of ${T}_{d}=150,100,50,25\,{\rm{K}}$ where the line ${T}_{d}=150\,{\rm{K}}$ represents the water snowline. Water ice sublimates in the region of ${T}_{d}\gt 150\,{\rm{K}}$.

Standard image High-resolution image
Figure 3.

Figure 3. Same as Figure 2, but for a disk around Herbig Ae/Be stars. Grain temperatures are higher due to the higher central star temperature, and the water snowline extends farther.

Standard image High-resolution image

The mean wavelength of the radiation field of a young star can be calculated by

Equation (4)

where the radiation intensity uλ (T) is described by the Planck function. This yields $\overline{\lambda }=1.33\,\mu {\rm{m}}$ for the T-Tauri disks and $\overline{\lambda }=0.53\,\mu {\rm{m}}$ for the Herbig Ae/Be disks. The radiation energy density is ${u}_{\mathrm{rad}}=\int {u}_{\lambda }d\lambda $.

The strength of the radiation field in the disk is characterized by a dimensionless parameter U = urad/uISRF where ${u}_{\mathrm{ISRF}}\,=8.64\times {10}^{-13}\,\mathrm{erg}\,{\mathrm{cm}}^{-3}$ is the energy density of the average interstellar radiation field (ISRF) in the solar neighborhood from Mathis et al. (1983). The local value of U in the disk can be approximately calculated using the grain temperature obtained from radmc-3d as follows:

Equation (5)

for silicate grains (see Draine 2011).

3. Rotational Disruption of Dust and Ice by Radiative Torques

3.1. Grain Suprathermal Rotation by RATs

When exposed to an anisotropic radiation field, a dust grain of irregular shape experiences RATs (Dolginov & Mitrofanov 1976) that can spin it up to suprathermal rotation (Draine & Weingartner 1996; Hoang & Lazarian 2009). Numerical calculations have revealed that RATs are weakly dependent on the dust composition (Lazarian & Hoang 2007; Herranen et al. 2019). Therefore, the following discussion is applied to both carbonaceous, silicate grains as well as carbon-silicate mixtures.

Let a be the effective grain size, which is defined as the radius of the equivalent sphere of the same volume. Following Hoang (2019), the maximum rotation rate of dust grains spun-up by a radiation field of anisotropy γ, mean wavelength $\bar{\lambda }$, and radiation strength U is given by

Equation (6)

for grains with $a\lesssim \bar{\lambda }/1.8$, and

Equation (7)

for grains with $a\gt \overline{\lambda }/1.8$. Here, ${n}_{1}={n}_{{\rm{H}}}/(10\,{\mathrm{cm}}^{-3})$, ${T}_{2}\,={T}_{\mathrm{gas}}/(100\,{\rm{K}})$, ${\bar{\lambda }}_{0.5}=\bar{\lambda }/(0.5\,\mu {\rm{m}})$, ${a}_{-5}=a/({10}^{-5}\,\mathrm{cm})$, and FIR is the dimensionless parameter describing the grain rotational damping by infrared emission that depends on nH, Tgas, and U (Draine & Lazarian 1998; Hoang et al. 2010). The rotation rate depends on the parameter $U/{n}_{{\rm{H}}}{T}_{\mathrm{gas}}^{1/2}$ and the damping by far-infrared emission FIR.

For convenience, let ${a}_{\mathrm{trans}}=\bar{\lambda }/1.8$, which denotes the grain size at which the RAT efficiency changes between the power-law and flat stages (see, e.g., Lazarian & Hoang 2007; Hoang et al. 2019) and ωRAT changes from Equation (6) to (7).

3.2. Rotational Disruption of Composite Grains

The centrifugal force applied to a spinning grain induces centrifugal stress S = ρω2a2/4, with ρ being the mass density of the grain; this stress tends to tear the grain apart. When the centrifugal stress exceeds the maximum tensile strength of the grain material, the grain is instantaneously disrupted into small fragments. This mechanism of dust destruction, developed by Hoang et al. (2019), is termed RATD.

The critical rotation rate required to disrupt the composite grain is described by

Equation (8)

where ${S}_{\max ,7}={S}_{\max }/({10}^{7}\,\mathrm{erg}\,{\mathrm{cm}}^{-3}$), with Smax being the maximum tensile strength of the grain and $\hat{\rho }=\rho /(3\,{\rm{g}}\,{\mathrm{cm}}^{-3})$ (Hoang et al. 2019).

The value of Smax depends on the composition, internal structures, and grain size. In PPDs, dust particles can have a large range of sizes, from nanoparticles to mm/cm sized pebbles, owing to the effect of grain coagulation and growth. To model RATD in the surface and intermediate layers of PPDs, we consider the size dependence of the tensile strength. Grains of size a > 0.1 μm can be considered as porous/composite. For these grains, we adopt a composite grain model as proposed by Mathis & Whiffen (1989), in which individual particles of silicate or carbonaceous materials are assumed to be compact and spherical of radius ap.

Following Greenberg et al. (1995), the maximum tensile strength of the grain is given by

Equation (9)

where $\bar{E}=\alpha {10}^{-3}\,\mathrm{eV}\,$ is the mean intermolecular interaction energy at the contact surface between two particles, h is the mean intermolecular distance, P is the porosity of the grain, and β is the mean number of contact points per particle between 1 and 10. For our estimates, we fix the porosity P = 0.2, as previously assumed for Planck data modeling (Guillet et al. 2018), and adopt the typical values α = 1, β = 5, h = 0.3 nm. Equation (9) yields ${S}_{\max }=3.2\times {10}^{5}\,\mathrm{erg}\,{\mathrm{cm}}^{-3}$ for ap = 10 nm.

Grains of a ≤ 0.05 μm are likely compact and have higher Smax than that of composite grains. We take the typical value ${S}_{\max }={10}^{9}\,\mathrm{erg}\,{\mathrm{cm}}^{-3}$ for these grains. For grain sizes in between ($0.1\,\mu {\rm{m}}\geqslant a\gt 0.05\,\mu {\rm{m}}$), we assume an intermediate value of ${S}_{\max }={10}^{7}\,\mathrm{erg}\,{\mathrm{cm}}^{-3}$.

Grain disruption sizes can be computed by solving the equation ωRAT = ωdisr. From Equation (8), one can see that ωdisr decreases as ∝a−1, whereas Equations (6) and (7) imply that ωRAT first increases with a until a = atrans and then decreases more rapidly than ωdisr as a goes beyond atrans. Hence, one can expect there are two intersections of ωRAT and ωdisr: one for a < atrans and one for a > atrans. Upon reaching the critical size determined by the first intersection of ωRAT and ωdisr, grains are disrupted. This value can be obtained from Equations (6) and (8), as given by

Equation (10)

for a ≲ atrans. The maximum size of grains that can still be disrupted is determined by the second intersection and can be estimated from Equations (7) and (8), as given by

Equation (11)

which depends on the local gas density and temperature due to gas damping. Under the effect of rotational disruption, all grains within the size range adisr ≤ a ≤ adisr,max would be disrupted.

In the absence of rotational damping, the time required to spin-up the grains of size adisr to ωdisr defines the disruption time:

Equation (12)

where U5 = U/105.

When rotational damping is present, solving the equation ω(t) = ωdisr yields the disruption timescale

Equation (13)

which is applicable for adisr,max > a > adisr. Note that, for a = [adisr, adisr,max], ${t}_{\mathrm{disr}}\to \infty $ because it takes t ≫ tdamp to reach ω = ωRAT. In strong radiation fields, when tdisr,0 ≪ τdamp, tdisr returns to tdisr,0.

3.3. Rotational Desorption of Ice Mantles

The rotational desorption mechanism of icy grain mantles was first studied by Hoang & Tram (2020), who found that the centrifugal stress induced by grain suprathermal rotation can desorb the entire ice mantle into tiny icy fragments, provided that the grain core is compact (see Figure 4).

Figure 4.

Figure 4. Schematic illustration of a rapidly spinning core-ice mantle grain of irregular shape, comprising a water-rich (dark blue) and a CO-rich (light blue) mantle layer. Complex molecules and PAHs are also present in the ice mantle.

Standard image High-resolution image

The grain model considered in Hoang & Tram (2020) is the one made of an amorphous silicate/carbonaceous core covered by a thick double-layer ice mantle (Oberg et al. 2010), where complex organic molecules are formed as illustrated in Figure 4. It is worth noting that the formation of ice mantles takes place on the grain surface, due to accretion of gas molecules in cold and dense regions of hydrogen density ${n}_{{\rm{H}}}=n({\rm{H}})+2n({{\rm{H}}}_{2})\sim {10}^{3}\mbox{--}{10}^{5}\,{\mathrm{cm}}^{-3}$ or the visual extinction AV > 3 (Whittet et al. 1983). Although icy grain mantles have nonspherical shapes, as demonstrated by dust polarization, let us assume that the grain shape can be described by an equivalent sphere of the same volume with effective radius a.

As discovered in Hoang & Tram (2020), the centrifugal stress applied to a spinning core-ice mantle grain tends to pull off the ice mantle from the grain core at sufficiently fast rotation. Upon reaching the critical disruption limit ωdisr, the centrifugal stress is equal to the tensile strength that the mantle exerts upon the grain core, which causes the ice mantle near the equator to detach. When the rotation rate increases beyond ωdisr such that the centrifugal stress exceeds the ice tensile strength that holds different parts of the mantle together, the mantle is disrupted into small fragments.

Let x0 be the distance from the core–mantle interface to the spinning axis. From Equation (7) in Hoang & Tram (2020), one obtains the centrifugal stress on the ice mantle:

Equation (14)

where ${\hat{\rho }}_{\mathrm{ice}}={\rho }_{\mathrm{ice}}/(1\,{\rm{g}}\,{\mathrm{cm}}^{-3})$ with ρice being the mass density of the ice mantle, which is ${\rho }_{\mathrm{ice}}\sim 1\,{\rm{g}}\,{\mathrm{cm}}^{-3}$ for pure ice (Hoang & Tram 2020), and ${\omega }_{10}=\omega /({10}^{10}\,\mathrm{rad}\,\,{{\rm{s}}}^{-1})$.

The critical rotational velocity for disruption is given by Sx = Smax:

Equation (15)

Similar to rotational disruption, one can obtain the desorption sizes of icy grains, as given by

Equation (16)

for $a\leqslant \overline{\lambda }/1.8$, and the maximum size of icy grains that can still be disrupted by centrifugal stress caused by RATs is described by

Equation (17)

In the absence of rotational damping, the characteristic timescale for rotational desorption can be estimated as:

Equation (18)

3.4. Ro-thermal Desorption of Molecules from Ice Mantles

In addition to rotational desorption, water and complex molecules can also be desorbed from icy grain mantles by the so-called rotational-thermal or ro-thermal desorption mechanism. Note that, in star-forming regions, molecules are assumed to be physically adsorbed to the ice mantle via the van der Waals force, also referred to as physisorption. In this mechanism, instead of evaporating from the disrupted icy fragments of the original mantle as studied in the previous subsection, molecules sublimate directly from the intact icy grain mantle with the support of centrifugal force. A detailed description and formulation of the ro-thermal desorption mechanism is presented in Hoang & Tung (2019). Here, we briefly describe the mechanism with regard to its application to the disks.

Let τsub,0 be the desorption rate of absorbed molecules with binding energy Eb from a grain at rest (ω = 0), which is heated to temperatures Td. Following Watson & Salpeter (1972), one has

Equation (19)

where ν is the characteristic vibration frequency of adsorbed molecules in the perpendicular direction to the lattice surface and given by

Equation (20)

with Ns being the surface density of binding sites (Tielens & Allamandola 1987). Typically, Ns ∼ 2 × 1015 site $\,{\mathrm{cm}}^{-2}$.

Following Hoang & Tung (2019), in the presence of grain suprathermal rotation, the centrifugal force acting on the absorbed molecule of mass m at distance $r\sin \theta $ from the spinning axis is

Equation (21)

where the unit vectors ($\hat{{\boldsymbol{x}}},\hat{{\boldsymbol{y}}}$) describe the plane perpendicular to the spinning axis, and ${{\boldsymbol{a}}}_{\mathrm{cen}}=-{\rm{\nabla }}{\phi }_{\mathrm{cen}}$ is the centrifugal acceleration. Here, the centrifugal potential

Equation (22)

which gives the average centrifugal potential

Equation (23)

where r ≡ a and $\langle {\sin }^{2}\theta \rangle =2/3$ is taken.

As a result, the effective binding energy of the absorbed molecule becomes

Equation (24)

The rate of ro-thermal desorption is then given by

Equation (25)

where the second exponential term describes the probability of desorption induced by centrifugal potential.

4. Numerical Results

4.1. Rotational Disruption of Dust Grains

We first show the results for disruption of dust grains in the surface and intermediate layers, which is a top-down mechanism to reproduce PAHs/nanoparticles as observed via mid-IR emission.

4.1.1. Disruption Sizes

To compute the disruption sizes of grains by RATD for a disk, we numerically solve the equation ωRAT = ωdisr. We consider the maximum grain size amax = 10 μm. We adopt γ = 0.7 for the anisotropy degree of the radiation field, which is appropriate for the surface and intermediate layers that are illuminated mostly by the stellar and ISRFs.

Figure 5 shows the obtained grain disruption sizes for a disk around T-Tauri stars with different values of Σ1 ranging from 1 to 1000. The gray shaded region indicates that no disruption occurs, due to the low radiation intensity and/or high gas density. Depending on the surface mass density of the disks, rotational disruption occurs in both the warm intermediate and surface layers with ${T}_{d}\gt 30\,{\rm{K}}$. Lower values of Σ1 result in an optically thin disk in which the disruption of grains is more prominent. In the disk surface, where the gas density ${n}_{{\rm{H}}}\lesssim {10}^{9}\,{\mathrm{cm}}^{-3}$ and grain temperature ${T}_{d}\gt 150\,{\rm{K}}$ (see Figure 2) for which grains could be spun-up to extremely fast rotation, rotational disruption is very efficient. The grain disruption size adisr decreases with the disk height z from ${a}_{\mathrm{trans}}\sim 0.74\,\mu {\rm{m}}$ to a minimum value ∼0.01 μm in the central region ($R\lesssim 0.2\,\mathrm{au}$), indicating that all grains with sizes larger than ∼10 nm would be disrupted in this region to form nanoparticles. In the warm layer, rotational disruption is still pronounced in the region where ${T}_{d}=50\mbox{--}150\,{\rm{K}}$ and ${n}_{{\rm{H}}}\lesssim {10}^{6}\,{\mathrm{cm}}^{-3}$, and near the upper boundary of the disk height, small grains of a ≳ 0.05 μm are already disrupted. For lower grain temperatures and higher gas density (i.e., ${T}_{d}=30\mbox{--}50\,{\rm{K}}$ and ${n}_{{\rm{H}}}\lesssim {10}^{2}\,{\mathrm{cm}}^{-3}$ for the case of ${{\rm{\Sigma }}}_{1}\,=1000\,{\rm{g}}\,{\mathrm{cm}}^{-2}$), rotational disruption becomes less efficient and only occurs at the surface layer with z/R = 0.4–0.5. Nevertheless, grains of a ≳ 0.1 μm are still destroyed under the effect of RATD at the height boundary. The maximum disruption size adisr,max increases rapidly with z and exceeds the threshold amax = 10 μm in the surface layer, which means only very large grains could survive. Overall, the closer to the central star, the smaller adisr and the greater adisr,max, which means that a broader range of grain sizes is disrupted by RATD.

Figure 5.

Figure 5. Disruption sizes of composite grains for a PPD around T-Tauri stars as functions of R and z/R, assuming different values of the surface mass density ${{\rm{\Sigma }}}_{1}=1\mbox{--}{10}^{3}\,{\rm{g}}\,{\mathrm{cm}}^{-2}$. White lines show the locations of the disk that have ${T}_{d}=150,100,50,25\,{\rm{K}}$. Black dotted lines mark the heights z = 1, 2, 3, 4, 5Hp.

Standard image High-resolution image

It is otherwise the same as Figure 5, but Figure 6 instead shows the results for the disk around Herbig Ae/Be stars. The active region of rotational disruption is larger due to higher temperatures in the surface layer. At $R\sim 30\,\mathrm{au}$, small grains of a ∼ 0.03 μm can still be disrupted compared to $R\sim 0.3\,\mathrm{au}$ in the case of T-Tauri disks. Beyond that radius, disruption size is ∼0.05 μm at the top of the scale height and remains relatively small up to the outer radius (0.01–0.05 μm).

Figure 6.

Figure 6. Same as Figure 5, but for a disk around Herbig Ae/Be stars. Rotational disruption is more efficient because of stronger radiation fields from the central star.

Standard image High-resolution image

To better understand rotational disruption in the different layers of the disk, in Figure 7, we show the disruption sizes as a function of the disk height z for the different disk radii, with ${{\rm{\Sigma }}}_{1}=1\,{\rm{g}}\,{\mathrm{cm}}^{-2}$. From Equations (10) and (11), one can see that adisr decreases when nH decreases and Td and U ∼ ${T}_{d}^{6}$ increases, whereas the opposite tendency is expected for adisr,max. As nH exponentially decreases and Td is significantly greater as one goes higher to the exposed surface layer (see Figures 2 and 3), adisr declines while adisr,max increases rapidly with z, making the range [adisr, adisr,max] larger. The discontinuity of the plots for adisr at a = 0.1 μm and a = 0.05 μm originates from the three separate values assumed for Smax for three ranges of grain sizes. Generally speaking, adisr mostly ceases at 0.05 μm, due to the large tensile strength ${S}_{\max }\sim {10}^{9}\,\mathrm{erg}\,{\mathrm{cm}}^{-3}$ of compact grain cores of $a\leqslant 0.05\,\mu {\rm{m}}$. Only near the central region (e.g., $R=1\,\mathrm{au}$) can disruption sizes be smaller than 0.05 μm. Note that the maximum grain size is set to be amax = 10 μm, which accounts for the stop at 10 μm of adisr,max. However, such large grains may not experience lower RATs, due to incoherent contributions of RATs from the different grain facets (see, e.g., Cho & Lazarian 2007). Since the calculations of RATs for very large grains of size $a\gt \bar{\lambda }/0.1$ are not yet available, due to computing limitations (see, e.g., Herranen et al. 2019), in the case of Herbig disks where $\bar{\lambda }/0.1=5.33\,\mu {\rm{m}}$, we plot a horizontal line where $a=\bar{\lambda }/0.1$, below which RATs have been previously calculated for irregular grains (Lazarian & Hoang 2007).

Figure 7.

Figure 7. Disruption sizes adisr and adisr,max as a function of the disk height z computed for different disk radii R, with ${{\rm{\Sigma }}}_{1}=1\,{\rm{g}}\,{\mathrm{cm}}^{-2}$. Upper and lower panels show the results for T-Tauri disks and Herbig Ae/Be disks, respectively. The value of adisr decreases and that of adisr,max increases with increasing z. Red horizontal line shows ${a}_{\mathrm{disr},\max }=\bar{\lambda }/0.1$.

Standard image High-resolution image

4.1.2. Rotational Disruption Time versus Grain–Grain Collision Destruction Time

We now calculate the time it takes for rotational disruption to disrupt the grains tdisr in the disks from Equation (12), using the numerical results for adisr and U from Section 4.1.1.

Other possible dust destruction mechanism is grain shattering caused by grain–grain collisions induced by radiation pressure (Hoang 2017). The mean time between two collisions defines the destruction time by grain–grain collisions

Equation (26)

where a single size a distribution with the gas-to-dust mass ratio Mg/d = 100 and the number density of dust grains ngr is assumed, and vgg is the relative velocity of grains. For the turbulence effect, one has vgg = αpcs, with the turbulence parameters α = 0.01 and p = 1/2 (see, e.g., Dullemond & Dominik 2004) and cs being the gas sound speed. Comparing Equation (26) with (12), one can see that grain–grain collisions require a much longer timescale to produce small grains compared to RATD.

In Figure 8, we compare the disruption time by RATD (left panel) with the shattering time by grain–grain collisions (right panel) for different grain sizes a for the T-Tauri disks, with ${{\rm{\Sigma }}}_{1}=1\,{\rm{g}}\,{\mathrm{cm}}^{-2}$. In the surface layer, where the gas and dust density are relatively low, the shattering time is extremely high, as τgg is most governed by nH. In contrast, the disruption time, which depends on disruption sizes, is much lower due to high radiation intensity from the central star that causes grains to disrupt more easily—and hence, adisr becomes smaller.

Figure 8.

Figure 8. Comparison of rotational disruption time tdisr (left panel) with shattering time by grain–grain collisions τgg (right panel) for grain size a = 1 μm and ${{\rm{\Sigma }}}_{1}=1\,{\rm{g}}\,{\mathrm{cm}}^{-2}$. The T-Tauri disk is considered. White solid line marks the boundary above which rotational disruption is effective.

Standard image High-resolution image

It is worth noting that, the higher the value of Σ1, the more massive the disk would be. A massive PPD, particularly around a T-Tauri star of 0.5M, is likely to be gravitationally unstable. In such a case, internal heating from gravitational instability–driven spirals and/or turbulence could dominate stellar radiation, giving rise to a disk model different from the passive irradiated disk adopted in Section 2. Furthermore, the Gaussian vertical profile of the gas described by Equation (1) assumes a vertically isothermal temperature structure, whereas the temperature profile from radmc-3d has a vertically stratified structure due to stellar radiation. As a result, the disk is not in hydrostatic equilibrium. However, the disruption time given by Equation (12) is smaller than the dynamical timescale, which is defined as the shortest timescale on which the disk structure can vary:

Equation (27)

where Ω is the Keplerian orbital period. The problem of disk instability is well-studied in the literature (see, e.g., Nakamoto & Nakagawa 1994; Umurhan & Shaviv 2005; Pfeil & Klahr 2019) and it is well-established that the thermal and viscous timescales are much larger than τD. Therefore, we expect that the disruption of grains would not be significantly affected in such events.

4.2. Rotational Desorption of Ice Mantles and the New Location of the Snowline

Now we move on to show our results for rotational disruption and desorption of ice mantles, which is relevant to the warm intermediate layer.

4.2.1. Desorption Sizes of Ice Mantles

We first consider a core-ice mantle grain model with a fixed core radius ac = 0.05 μm and a mantle thickness Δam that can vary such that the grain size is a = ac + Δam. Similar to disruption sizes, numerical calculations for desorption sizes are performed with conservative values for the tensile strength of the grain core ${S}_{\max ,\mathrm{core}}={10}^{9}\,\mathrm{erg}\,{\mathrm{cm}}^{-3}$ and of the ice mantle ${S}_{\max ,\mathrm{mantle}}={10}^{7}\,\mathrm{erg}\,{\mathrm{cm}}^{-3}$, which is comparable to the tensile strength of bulk ice (see, e.g., Litwin et al. 2012).

The results for the T-Tauri disk are plotted in Figure 9. The snowline (${T}_{d}=150\,{\rm{K}}$) obtained from radmc-3d is shown for comparison. The similar variation features as rotational disruption can be observed here. Rotational desorption of ice mantles also occurs in the warm intermediate and surface layers with ${T}_{d}\gt 30\,{\rm{K}}$, where grains are spun-up to suprathermal rotation and the ice mantles are disrupted due to centrifugal stress. The desorption size adesp is smaller and adesp,max is greater near the central star because of the higher radiation intensity and lower gas number density. In the surface layer where ${T}_{d}\gt 150\,{\rm{K}}$, icy grains of sizes larger than the grain core ac = 0.05 μm are mostly desorbed. The intermediate layer where ${T}_{d}=30\mbox{--}150\,{\rm{K}}$ also witnesses great efficiency of rotational desorption, with adesp drastically increasing from atrans = 0.74 μm to ∼ac = 0.05 μm as R declines and z increases. Interestingly, rotational desorption takes place in the region beyond the water snowline (around $3\,\mathrm{au}$ from the star) as well. Consequently, the ice mantles of dust grains in this region are already disrupted into smaller fragments, giving rise to the subsequent evaporation into water vapor. As a result, rotational desorption can destroy the ice mantles of grains beyond the snowline—and therefore push down the snowline in the vertical direction to the boundary beyond which the ice mantles start to be disrupted under the RATD effect.

Figure 9.

Figure 9. Desorption sizes of ice mantles for a PPD around T-Tauri stars as a function of R and z/R, assuming a fixed core radius ac = 0.05 μm and the varying mantle thickness. White dashed and dotted lines illustrate the region where ${\tau }_{\mathrm{sub},\mathrm{rot}}^{-1}({T}_{d})$ = ${\tau }_{\mathrm{sub},0}^{-1}({T}_{\mathrm{sub}})$ for different grain sizes a, assuming water ice of ${T}_{\mathrm{sub}}=150\,{\rm{K}}$. Different values of the surface mass density Σ1 are considered.

Standard image High-resolution image

Furthermore, to demonstrate the effect of ro-thermal desorption on the release of COMs beyond the water snowline, we plot also in Figure 9 the region where ${\tau }_{\mathrm{sub},\mathrm{rot}}^{-1}({T}_{d})\,={\tau }_{\mathrm{sub},0}^{-1}({T}_{\mathrm{sub}})$ for different grain sizes a, where ${T}_{\mathrm{sub}}=150\,{\rm{K}}$ is taken to be the sublimation temperature and ${E}_{b}=5700\,{\rm{K}}$ the binding energy of water ice (Garrod 2013). In the surface layer, the snowlines in the cases with and without the rotation of grains become separated due to high Td and low nH increasing the rotation rate and thus decreasing the binding energy of molecules. As a result, the region that has a desorption rate equal to that of water ice by thermal desorption is extended further from the central star. The effect is more significant for larger grain sizes. Our model of ro-thermal desorption is not limited only to the water snowline. Plugging in the analogous parameters for a CO molecule, i.e., ${E}_{b}=1150\,{\rm{K}}$ (Garrod 2013) and ${T}_{\mathrm{sub}}=25\,{\rm{K}}$ (Mumma et al. 1993), one can see that the rotation of grains induced by RAT also increases the desorption rate of CO ice. Thus, the CO iceline under the framework of ro-thermal desorption would also be extended compared to the classical one. It is worth noting that the effect of rotational desorption is more efficient than that of ro-thermal desorption, i.e., rotational desorption can occur at more distant regions.

Figure 10 shows results similar to those of Figure 9, but for the disk around Herbig Ae/Be stars. The active region of rotational desorption is larger due to higher temperatures in the surface layer. Rotational desorption is also very efficient in the region beyond the snowline, which is much further from the star in this case (around $40\,\mathrm{au}$ from the central star). Up to $\sim 200\,\mathrm{au}$, desorption sizes still can reach $0.05\,\mu {\rm{m}}$ in the upper layer. At the outer radius $R=300\,\mathrm{au}$, the desorption of icy grains has not eased for high values of Σ1 (e.g., 100 and $1000\,{\rm{g}}\,{\mathrm{cm}}^{-2}$) as in the case of T-Tauri disks. Thus, the snowline is pushed down further by rotational desorption in this case.

Figure 10.

Figure 10. Same as Figure 9, but for a disk around Herbig Ae/Be stars. Due to higher dust temperatures induced by a stronger radiation field, rotational desorption takes place in a broader region.

Standard image High-resolution image

In Figure 11, we plot the results for ro-thermal desorption with ${{\rm{\Sigma }}}_{1}=1\,{\rm{g}}\,{\mathrm{cm}}^{-2}$ in regular R − z coordinates, for a better illustration of this effect on the snowline location.

Figure 11.

Figure 11. Illustration of the water snowline for disks around T-Tauri (left panel) and Herbig stars (right panel), due to ro-thermal desorption. Ro-thermal desorption of ice mantles extends the original snowline, depending on the radius of icy grains.

Standard image High-resolution image

Figure 12 shows the desorption sizes of icy grains as a function of the disk height z at the different positions along the disk radius, with ${{\rm{\Sigma }}}_{1}=1\,{\rm{g}}\,{\mathrm{cm}}^{-2}$. One can see that adesp decreases exponentially with z to the lower boundary  0.05 μm, while adesp,max increases exponentially until reaching the maximum value of ∼10 μm due to the decrease of nH and the increase of Td, similar to rotational disruption.

Figure 12.

Figure 12. Desorption sizes adesp and adesp,max as a function of the disk height z computed for different radii R. Upper and lower panels show the results for T-Tauri disks and Herbig Ae/Be disks, respectively. The value of adesp decreases and that of adesp,max increases with increasing z. Red horizontal line shows ${a}_{\mathrm{disr},\max }=\bar{\lambda }/0.1$.

Standard image High-resolution image

To quantify the effect of rotational desorption on the presence of ice in the disks, we calculate the total surface mass density of water ice disrupted by rotational desorption

Equation (28)

where ${V}_{\mathrm{ice}}=4\pi ({a}^{3}-{a}_{{\rm{c}}}^{3})/3$ is the volume of the ice mantle of grain of size a with the core radius ac, and the power-law distribution dngr(a)/da = Caq is again assumed for icy grains. We consider a minimum grain size amin = 0.05 μm and a maximum grain size amax = 0.10 μm. The normalization coefficient C is calculated assuming the ratio of gas-to-dust mass to be 100. For thermal sublimation, we assume that the ice mantles of all grains with ${T}_{d}\gt 150\,{\rm{K}}$ are removed and compute the amount of sublimated ice ΔΣice,sub. Results for ΔΣice,disr and ΔΣice,sub are shown in Figure 13. Prior to the snowline (at $R\sim 3\,\mathrm{au}$ for the T-Tauri disks and $R\sim 40\,\mathrm{au}$ for the Herbig Ae/Be disks), the amount of water ice sublimated through thermal sublimation is higher than that of ice disrupted by rotational desorption. This is expected, given that ice is believed to be present only in the region beyond the snowline where grain temperature is lower than the sublimation threshold ${T}_{d}\sim 150\,{\rm{K}}$ of H2O. Therefore, according to our assumption, the mantles of all grains located before the snowline from the disk midplane to the surface are destroyed by thermal sublimation, whereas rotational does not occur in the disk interior and thus can affect a smaller surface mass density of water ice. However, beyond the snowline, rotational desorption continues to happen in the upper layer, and ice can still be removed in this region. The amount of disrupted ice is higher for smaller maximum grain sizes, for which more dust grains in the range size $[{a}_{\min },{a}_{\max }]$ are under the effect of RATD.

Figure 13.

Figure 13. Surface mass density ΔΣice of water ice disrupted by rotational desorption (solid lines) and of water ice sublimated by thermal sublimation (dashed lines). Thermal sublimation is the main mechanism to desorb water prior to the snowline, but rotational desorption takes over beyond the snowline. Left and right panels show the results for T-Tauri disks and Herbig Ae/Be disks, respectively.

Standard image High-resolution image

4.2.2. Rotational Desorption Time versus Sublimation Time

Now let us compare the thermal sublimation time to the time it takes for rotational desorption to disrupt the ice mantles.

Following Hoang & Tram (2020), the sublimation time of the ice mantle of thickness Δam is given by

Equation (29)

where da/dt ∼ l/τsub, with l being the thickness of ice monolayer, is the rate of decrease in the mantle thickness due to thermal sublimation.

We compute tsub for desorption of water molecules with the binding energy ${E}_{b}=5700\,{\rm{K}}$, using the dust temperature from radmc-3d and the desorption time for the disks in the region where rotational desorption of ice mantles takes place from Equation (18), using numerical results for adesp from Section 4.2. A comparison is provided in Figure 14 for grain size a = 1 μm and ${{\rm{\Sigma }}}_{1}=1\,{\rm{g}}\,{\mathrm{cm}}^{-2}$. Rotational desorption of ice mantles is much faster than thermal sublimation in the surface and intermediate layers. The difference could be of order 10 in the inner central region.

Figure 14.

Figure 14. Comparison of desorption time of water ice tdesp (left panel) with the classical sublimation time (right panel) for a disk around T-Tauri star. Grain size a = 1 μm and ${{\rm{\Sigma }}}_{1}=1\,{\rm{g}}\,{\mathrm{cm}}^{-2}$ are considered. Rotational desorption is much faster than sublimation in the active region (above the white line, right panel).

Standard image High-resolution image

4.3. Ro-thermal Desorption of Molecules from Ice Mantles

4.3.1. Ro-thermal Desorption of Molecules

We have seen that rotational desorption and ro-thermal desorption of ice mantles are both very efficient mechanisms to desorb COMs. As shown in Figures 9 and 10, for the grain model with a fixed core radius ac = 0.05 μm and a varying mantle thickness, the region where rotational desorption is important appears to be larger than that of ro-thermal desorption. However, for a grain model of thin ice mantle, the resulting tensile stress by suprathermal rotation may be insufficient to overcome the adhesive energy between the mantle and the grain core surface. Therefore, it is hard to disrupt the entire ice mantle via RATD. In this case, the tensile strength is replaced by the adhesive strength, which depends on the mechanical properties of the surface and grain temperature. The adhesive strength is low for a clean surface, but it can reach $\sim {10}^{9}\,\mathrm{erg}\,{\mathrm{cm}}^{-3}$ for some rough surfaces (Work & Lian 2018). As shown in Hoang & Tung (2019), ro-thermal desorption of individual molecules can occur before the disruption of ice mantle if the ice mantle thickness is below 100 monolayers of water ice (i.e., Δam < 200 Å).

We thus consider a core-ice mantle model in which the mantle thickness Δam is fixed to a certain value while the core radius is varied. Results with ${{\rm{\Sigma }}}_{1}=1\,{\rm{g}}\,{\mathrm{cm}}^{-2}$ are shown in 15. For Δam = 100 Å, the effect of rotational desorption of ice mantles becomes less significant than ro-thermal desorption for a = 0.5 μm in the case of T-Tauri disks and for a = 0.2 μm in the case of Herbig Ae/Be disks. For thinner mantles, the region where ice mantles are disrupted is narrower, and the active region of ro-thermal desorption is sufficiently broader for larger grain sizes. In that case, ro-thermal desorption takes over rotational desorption of ice mantles to release water and COMs into the gas phase. Furthermore, for this case of a thin mantle, the effect of rotational desorption is comparable to ro-thermal desorption, but less efficient than for the case of a thick ice mantle (see Figure 14).

Figure 15.

Figure 15. Desorption sizes of ice mantles adesp as a function of R and z/R for T-Tauri disks (left panels) and Herbig Ae/Be disks (right panels) for the model with a fixed mantle thickness Δam and a varying core radius. Adhesive strength ${S}_{\max ,\mathrm{mantle}}={10}^{9}\,\mathrm{erg}\,{\mathrm{cm}}^{-3}$ is assumed for the ice mantle. Different values of Δam are considered. Rotational desorption is less efficient for a thinner ice mantle.

Standard image High-resolution image

4.3.2. Ro-thermal Desorption of PAHs/Nanoparticles from the Ice Mantle

Ro-thermal desorption is found to be an efficient mechanism to desorb PAHs that are weakly bound to the ice mantle via the van der Waals force (Hoang & Tung 2019). Since ro-thermal desorption requires lower radiation strength to desorb than rotational desorption, one can describe the efficiency of ro-thermal desorption by considering the ejection threshold. Following Hoang & Tung (2019), one obtains the ejection threshold of PAHs:

Equation (30)

where the respective binding energies of benzene C6H6 and naphthalene (C10H8) to ice are ${E}_{b}/k\sim 4000\,{\rm{K}}$ and $6000\,{\rm{K}}$ (see Tables 4 and 5 in Michoulier et al. 2018).

The ejection radiation strength is then

Equation (31)

for a ≲ atrans, and

Equation (32)

for a > atrans. Clearly, the ejection threshold is much lower than that of water and COMs. Therefore, the ro-thermal desorption is efficient for desorption of PAHs in star-forming regions.

5. Effect of Rotational Disruption on Dust Opacity

Assuming no disruption, the dust opacity, defined as the total absorption cross section per unit of dust mass, is given by

Equation (33)

where Qabs,sca(a, λ) is the absorption/scattering efficiency for a grain of radius a at wavelength λ, dngr/da is the grain size distribution, X(a) = πa2Qabs,sca(a, λ)(dngr/da) and Y(a) = (4πρa3/3)(dngr/da). Here, we also assume a power-law distribution grain size dngr/da = CnHaqda with q = 3.5 and the ratio of gas-to-dust mass to be 100, as previously done in Section 4.2.1 but for composite grains of mass density $\rho =3\,{\rm{g}}\,{\mathrm{cm}}^{-3}$.

Due to rotational disruption, which redistributes all grains of size $a=[{a}_{\mathrm{disr}}-{a}_{\mathrm{disr},\max }]$ to smaller sizes (a < adisr), the dust opacity is modified to

Equation (34)

where q' indicates that a new size distribution is assigned for the resulting grains and can be determined from the conservation of dust mass:

Equation (35)

where V(a) = 4πρa3/3 is the grain volume. Here, we assume amin = 0.01 μm and amax = 10 μm.

From the numerical results in Section 4, we study how the dust opacity is modified by RATD in each cell, using the absorption/scattering cross section for amorphous silicate grains as previously done in Section 2.3. Table 2 shows the obtained values of q' at several cells in the disk, along with the locations of the cells and the disruption sizes calculated for those cells. In Figure 16, we show the scattering opacity κsca at a position in the disk surface for two types of PPDs, with ${{\rm{\Sigma }}}_{1}=1\,{\rm{g}}\,{\mathrm{cm}}^{-2}$. The lower and upper boundaries for disruption sizes of grains by RATD in the numerical calculations without taking into account ${a}_{\mathrm{disr},\max }=\overline{\lambda }/0.1$ are shown for comparison in the case of Herbig disks. One can see that κsca decreases substantially at optical to mid-infrared ($\lambda \sim 0.5\mbox{--}50\,\mu {\rm{m}})$ due to the disruption of dust grains. This arises from the fact that dust grains efficiently scatter photons with a wavelength comparable to the grain size, such that when those grains are removed by RATD, the scattering opacity is decreased accordingly. Note that, in the case of Herbig disks, the upper boundary of disruption size at the chosen point exceeds $\overline{\lambda }/0.1=5.33\,\mu {\rm{m}}$. As a result, there is a considerable difference between the results with and without considering ${a}_{\mathrm{disr},\max }=\overline{\lambda }/0.1$.

Figure 16.

Figure 16. Scattering opacity κsca for the different maximum grain sizes at selected positions in the surface layer of the disk as a function of the wavelength λ with (solid lines) and without (dashed lines) the presence of disruption. Dotted line shows the results when considering ${a}_{\mathrm{disr},\max }=\overline{\lambda }/0.1$. Left and right panels show the results for T-Tauri disks and Herbig Ae/Be disks, respectively.

Standard image High-resolution image

Table 2.  Modified Slopes of Grain Size Distribution by RATD

Disk Type R z adisr adisr,max q q'
  (au) (au) (μm) (μm)    
T-Tauri 5.31 0.50 NDa ND 3.5 3.5
  5.31 1.05 0.6339 1.2218 3.5 3.53
  5.31 2.01 0.0496 10.000 3.5 3.75
  19.94 4.98 0.3488 9.8517 3.5 3.66
Herbig Ae/Be 10.33 1.02 ND ND 3.5 3.5
  10.33 4.00 0.0270 10.000 3.5 3.79
  52.05 9.85 0.1609 2.4039 3.5 3.63
  102.09 20.20 0.1271 5.4929 3.5 3.68

Note.

aNo disruption.

Download table as:  ASCIITypeset image

Figure 17 shows the absorption opacity κabs at a chosen disk radius and height in the surface layer of the disks around T-Tauri and Herbig Ae/Be stars for the cases with and without disruption. Though the difference is less pronounced than in the case of κabs, there is a considerable decrease of κabs in the wavelengths between ∼0.4 μm and ∼8 μm, which corresponds to the optical and near/mid-IR regions. This can be expected from the destruction of dust grains in the surface layer, where most of the mid-IR emissions of the disks come from.

Figure 17.

Figure 17. Same as Figure 16, but for the absorption opacity κabs. Rotational disruption reduces κabs less effectively than ksca.

Standard image High-resolution image

In Figure 18, we keep z/R constant and plot the opacities κabs and κsca versus the disk radius for the T-Tauri disks at λ = 7 μm. Doing so, we are able to see how κabs and κsca vary with R. For higher values of z/R, the inner region experiences a significant decrease in opacity, due to the very strong effect of rotational disruption, while for lower values, the difference is not pronounced, due to the absence of RATD in the interior. The disruption of large grains will consequently decrease the intensity of scattered light in NIR-MIR wavelengths from the surface layer of the disks. Many observations of PPDs have detected light scattered by dust grains on the surface layer, e.g., VLT/SPHERE NIR observations of the RY Lup disk around a T-Tauri star by Langlois et al. (2018), and Gemini Planet Imager observation of Herbig Ae/Be stars HD 150193, HD 163296, and HD 169142 by Monnier et al. (2017). Stolker et al. (2016) performed the scattered light mapping on the observational data from VLT/SPHERE and VLT/NACO of the disk around HD 100546 star and concluded that scattering opacity in the surface layer is mostly due to large, aggregate dust grains. Therefore, we expect that disruption of these grains by RATD, which reduces κsca, will have an important effect on NIR-MIR observations.

Figure 18.

Figure 18. Scattering opacity κsca (upper panels) and absorption opacity κabs (lower panels) as a function of the disk radius R for different values of z/R with (solid lines) and without (dashed black line) the presence of rotational disruption. Dotted lines show the results for ${a}_{\mathrm{disr},\max }=\overline{\lambda }/0.1$. Left and right panels show the results for T-Tauri and Herbig Ae/Be disks, respectively.

Standard image High-resolution image

6. Discussion

6.1. RATD as a Top-down Mechanism to form PAHs/Nanoparticles in the Disk Surface Layers

It is difficult to explain the widespread presence of PAHs/nanoparticles in the surface layer of PPDs around Herbig Ae/Be stars, even in the inner gap (e.g., Boutéraon et al. 2019), because PAHs/nanoparticles in the surface layer are expected to be destroyed by EUV photons and the X-ray component of the star's radiation spectrum (Siebenmorgen & Krügel 2010). To explain the observations of PAHs in the disk, Siebenmorgen & Krügel (2010) suggested that PAHs are transported from the disk interior to the surface via turbulent mixing in the vertical direction, which requires the presence of PAHs in the disk interior. This scenario is difficult to reconcile with the fact that PAHs are expected to be depleted due to condensation into the ice mantle of dust grains in cold dense clouds (Bernstein et al. 1999; Cuylle et al. 2014; Cook et al. 2015). In this paper, we propose a top-down mechanism in which PAHs/nanoparticles are produced from the disruption of dust grains when they are being transported from the disk interior to the surface layers.

Our detailed modeling shows that large grains can be disrupted into nanoparticles in a vast area of the surface layer by the RATD mechanism. Moreover, PAHs/nanoparticles, which are frozen in the icy grain mantles, can desorb from the ice mantle via the ro-thermal desorption mechanism (Hoang & Tung 2019). Thus, if grains are small enough (a ≲ 1 μm) to be mixed with the gas by turbulence, they can escape grain settling and coagulation in the disk mid-plane and are transported to the surface layer (Dubrulle et al. 1995; Dullemond & Dominik 2004; Fromang & Nelson 2006, 2009). When exposed to stellar radiation, large grains within the range size [adisr, adisr,max] could be disrupted by RATD into PAHs/nanoparticles (see Figure 7). Furthermore, our results suggest an increase in the relative abundance of nanoparticles that corresponds with decreasing the distance to the star. Therefore, in the framework of rotational disruption and desorption, PAHs and nanoparticles would be continuously replenished in the surface layer, with the support of turbulent transport. This can resolve the longstanding mystery regarding the ubiquitous presence of PAHs/nanoparticles in the hot surface layer of PPDs.

Finally, spectroscopic and polarimetric measurements of the debris disk HD 32297 from Bhowmik et al. (2019) found an unusual blue NIR spectrum that corresponds to the presence of small grains of size much smaller than the bow-out limit due to radiation pressure (Wyatt et al. 1999). The authors explained this feature by a combination of several physical processes, including steady-state collisional evolution and the chain reaction of dust grain collisions triggered by the initial breakup of a large planetesimal near the central star, which Grigorieva et al. (2007) named the "avalanche" process. However, this process needs a two-belt structure for the breakups of massive planetesimals and the development and propagation of avalanche chain reaction (Thebault & Kral 2018).

6.2. Implications for the Detection of Gas-phase Water and Complex Molecules in PPDs

Our numerical results for desorption sizes presented in Section 4.2 indicate that rotational desorption is very efficient in the disk surface. The ice mantles of icy grains of size in the range adesp < a < adesp,max are disrupted into small fragments. These tiny fragments can then be transiently heated to evaporation temperatures, inducing transient release of molecules, as theoretically demonstrated in Hoang & Tram (2020). Therefore, water ice and complex molecules frozen in the ice mantle can be released into the gas phase even at temperatures lower than their sublimation threshold. In particular, we predict that gas-phase water and COMs could be detected at locations well beyond the snowline of R > 3–100 au (see Figures 9 and 5).

To date, several complex molecules have been detected in PPDs, including CH3OH and (HCOOH) in the TW Hydrae disk by Walsh et al. (2016) and Favre et al. (2018), and CH3CN by Öberg et al. (2015). Recently, ALMA observations of the DG Tau disk by Podio et al. (2019) showed a lack of H2CO emission in the warm inner disk, while the emission peak is located beyond the CO iceline ($\sim 40\,\mathrm{au}$). Interestingly, these molecules are detected far beyond the snowline at a distance above $10\,\mathrm{au}$. This cannot be explained by thermal sublimation, but rather is achieved via ro-thermal desorption as well as rotational desorption as shown in Figures 9 and 5.

We note that nonthermal desorption mechanisms such as cosmic-ray-induced desorption and UV photodesorption are not ruled out in PPDs (see, e.g., Öberg et al. 2015). Nevertheless, Walsh et al. (2010) carried out physical and chemical modeling of the PPD around the T-Tauri stars and found that cosmic-ray-induced desorption has little effect on the gas-phase abundances in the disk.

6.3. Implications for Formation of Comets and Asteroids

The water snowline characterized by the water sublimation temperature of ${T}_{d}\sim 150\,{\rm{K}}$ divides the inner region of rocky planets from the outer region of gas giant planets. Nevertheless, the precise location of the snowline is a longstanding problem in planetary science (Lecar et al. 2006; Min et al. 2011). In this paper, we studied the rotational desorption of ice mantles from the grain surface in PPDs.

Using the standard model of PPDs, we show that ice mantles from micron-sized grains (a < 100 μm) are disrupted and identify the location of snowline in the presence of rotational desorption. We find that the water snowline is extended outward. Figures 9 and 10 show that rotational desorption takes place in the region beyond the snowline—and thus can destroy the ice mantles of grains in this region. Indeed, observational detection of water ice from edge-on disks is reported in Terada et al. (2007) using water ice absorption features (Inoue et al. 2008). Future telescopes such as SPHEREx, GMT, the James Webb Space Telescope, SUBARU, and WFIRST would be useful to test the snowline. The rotational desorption of ice can have important implications for formation of planets, comets, and asteroids. Indeed, the total mass of rotationally disrupted ice is ${M}_{\mathrm{ice},\mathrm{disr}}$ = ${\int }_{{R}_{\mathrm{snowline}}}^{{R}_{\mathrm{out}}}2\pi r{\rm{\Delta }}{{\rm{\Sigma }}}_{\mathrm{ice},\mathrm{disr}}(r){dr}$${10}^{23}\mbox{--}{10}^{24}\,{\rm{g}}$ for T-Tauri disks and $\sim {10}^{25}\,{\rm{g}}$ for Herbig Ae/Be disks. As a result, the number of comets formed in the solar nebula would be significantly lower than previous estimates. On the other hand, the number of asteroids would be accordingly higher.

6.4. Rotational Disruption and the Problem of Carbon Deficit of the Earth and Rocky Bodies

Various studies indicate that the C/Si abundance ratio in the Earth and rocky bodies in the inner solar system is several orders of magnitudes lower than the primordial interstellar ratio (see, e.g., Lee et al. 2010; Bergin et al. 2015; Anderson et al. 2017). Here, we suggest that rotational disruption can explain the carbon deficit if there exist two separate populations of carbonaceous and silicate grains in PPDs. Note that the two separate populations originate naturally from the rotational disruption of large aggregates by RATD if these original grains are a carbon-silicate mixture as expected from grain coagulation in dense regions.

Indeed, our present studies imply that carbonaceous nanoparticles can be reproduced by the disruption of large carbonaceous grains that are being transported from the disk interior to the surface layer (see Figures 5 and 6). When being exposed to harsh UV radiation, PAHs/carbonaceous nanoparticles are destroyed, converting solid carbon into the gas phase. The disruption of silicate grains by RATD is expected to be less efficient because of the effect of grain alignment. Indeed, silicate grains, which are paramagnetic material, are aligned by RATs along the ambient magnetic fields, a phenomenon referred to as B-RAT alignment; see Andersson et al. (2015) and Lazarian et al. (2015) for reviews. On the other hand, carbonaceous grains are diamagnetic material; as such, they are unlikely to be aligned with the magnetic field, rather than the radiation direction (Lazarian & Hoang 2007). If the stellar radiation makes an angle ψ with the ambient magnetic field, the rotation rate driven by RATs for silicate grains is smaller than that of carbon grains by a factor $\cos \psi $ (Hoang & Lazarian 2009, 2014) if the Larmor precession rate (${\tau }_{\mathrm{Lar}}^{-1}$) exceeds the precession rate around the radiation (τk−1; Hoang & Lazarian 2016). The B-RAT alignment can occur closer to the star if silicate grains contain iron inclusions (Hoang & Lazarian 2016). The magnetic field geometry in PPDs is uncertain, but for an hourglass-shaped magnetic field, one expects that the stellar radiation forms a large angle with the magnetic field in the surface layers, i.e., $\cos \psi \lt 1$. As a result, RATD is, in general, more efficient for carbonaceous grains than for silicate grains, thus enhancing the conversion of large carbons into nanoparticles as well as the depletion of carbon depletion. A detailed study will be presented elsewhere.

7. Summary

We have studied the disruption of dust and of ice mantles in PPDs using the newly discovered rotational disruption mechanism based on centrifugal force within rapidly spinning dust grains induced by radiative torques. The principal results are summarized as follows.

  • 1.  
    We performed two-dimensional modeling of rotational disruption of composite grains by radiative torques, namely the RATD effect, for the PPDs around T-Tauri and Herbig Ae/Be stars. We find that the RATD effect is efficient in disrupting large grains into nanoparticles. The disruption time is found to be much shorter than the grain–grain collision time in the warm and hot surface layers.
  • 2.  
    We also modeled rotational disruption and desorption of ice mantles. We found that the ice mantles can be disrupted even in the warm intermediate layer with grain temperatures ${T}_{d}\sim 30\mbox{--}150\,{\rm{K}}$. We also found that the water and COMs can be desorbed rapidly as a result of ro-thermal desorption from tiny fragments produced by RATD.
  • 3.  
    We quantified the effect of RATD on the dust opacity and found that rotational disruption of grains of sizes a ∼ 0.1–10 μm significantly reduces the scattering opacity in optical- MIR wavelengths. As a result, observations of scattered light from the surface layers of PPDs must take into account the effect of rotational disruption.
  • 4.  
    Our results suggest that rotational disruption of ice mantles and fluffy grains could reproduce PAHs/nanoparticles, which are usually observed in the hot surface layers of PPDs. The subsequent destruction of PAHs by extreme UV photons would induce the deficit of carbons in the PPDs, resolving the problem of C deficit in terrestrial planets.
  • 5.  
    We showed that the water snowline defined by rotational desorption of ice mantles is more extended than the classical snowline predicted by water evaporation temperature. The amount of water ice destroyed by rotational desorption is significant up to a distance of $100\,\mathrm{au}$, which decreases the number of comets and increases the number of asteroids.
  • 6.  
    Our detailed modeling shows that rotational desorption and ro-thermal desorption of ice mantles are still active beyond the snowline. Thus, future modeling of chemistry in disks should take into account the effect of rotational disruption and desorption of dust and ice.
  • 7.  
    We suggest that rotational disruption by RATs may resolve the carbon deficit problem in the Earth and rocky bodies in the inner Solar System because carbonaceous grains may be disrupted more efficiently than silicates, due to them having a rotation rate higher than that of silicate grains.

We are grateful to the anonymous referee for helpful comments that improved our calculations and the presentation of the manuscript. N.D.T thanks Le Ngoc Tram, R.R. Rafikov, and C.P. Dullemond for helpful discussions during the early stage of this work. This research was supported by National Research Foundation of Korea (NRF) grants funded by the Korean government (MSIT) through the Basic Science Research Program (2017R1D1A1B03035359) and Mid-career Research Program (2019R1A2C1087045).

Footnotes

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