Estimate on Dust Scale Height from the ALMA Dust Continuum Image of the HD 163296 Protoplanetary Disk

and

Published 2021 May 17 © 2021. The American Astronomical Society. All rights reserved.
, , Citation Kiyoaki Doi and Akimasa Kataoka 2021 ApJ 912 164 DOI 10.3847/1538-4357/abe5a6

Download Article PDF
DownloadArticle ePub

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

0004-637X/912/2/164

Abstract

We aim at estimating the dust scale height of protoplanetary disks from millimeter continuum observations. First, we present a general expression of intensity of a ring in a protoplanetary disk and show that we can constrain the dust scale height by the azimuthal intensity variation. Then, we apply the presented methodology to the two distinct rings at 68 au and at 100 au of the protoplanetary disk around HD 163296. We constrain the dust scale height by comparing the high-resolution millimeter dust continuum image obtained in the Disk Substructures at High Angular Resolution Project (DSHARP) with radiative transfer simulations using RADMC-3D. We find that hd/hg > 0.84 at the inner ring and hd/hg < 0.11 at the outer ring with 3σ uncertainties, where hd is the dust scale height and hg is the gas scale height. This indicates that the dust is flared at the inner ring and settled at the outer ring. We further constrain the ratio of the turbulence parameter α to the gas-to-dust-coupling parameter St from the derived dust scale height; α/St > 2.4 at the inner ring, and α/St < $1.1\times {10}^{-2}$ at the outer ring. This result shows that the turbulence is stronger or the dust is smaller at the inner ring than at the outer ring.

Export citation and abstract BibTeX RIS

1. Introduction

The Atacama Large Millimeter/submillimeter Array (ALMA) has changed the picture of protoplanetary disks, which are the fields of planet formation. High spatial resolution observations with ALMA have revealed that many protoplanetary disks have structures in the millimeter dust continuum such as spirals (Pérez et al. 2016), crescents (van der Marel et al. 2013; Fukagawa et al. 2013; Casassus et al. 2015; Isella et al. 2018), and rings (ALMA Partnership et al. 2015; Fedele et al. 2017, 2018; van der Plas et al. 2017; Clarke et al. 2018; Dipierro et al. 2018; Andrews et al. 2018). The mechanism of dust ring formation is still under discussion. There are several possible scenarios for the dust ring formation: dust accumulation by gas gaps induced by planets (e.g., Lin & Papaloizou 1979; Goldreich & Tremaine 1980; Zhu et al. 2012; Pinilla et al. 2012a, 2012b), enhanced dust fragmentation by dust sintering near snowlines (Okuzumi et al. 2016), dust accumulation at the outer edge of the dead zone (Gressel et al. 2015; Flock et al. 2015; Ueda et al. 2019), and the secular gravitational instability (SGI; Ward 2000; Youdin 2011; Michikoshi et al. 2012; Takahashi & Inutsuka 2014; Tominaga et al. 2019).

The physical parameters of protoplanetary disks have been estimated by various methods. The dust radius has been estimated by the spectral index of the dust emission (e.g., Testi et al. 2014), and the gas surface density has been estimated by the line emission of gas (e.g., Isella et al. 2016). The turbulence α is an important parameter: it controls the accretion of disks (e.g., Shakura & Sunyaev 1973; Lynden-Bell & Pringle 1974; Hartmann et al. 1998), growth and fragmentation of the dust (e.g., Blum & Münch 1993; Blum & Wurm 2008; Brauer et al. 2008; Wada et al. 2013), and mixing of the dust and the gas (e.g., Dubrulle et al. 2005; Youdin & Lithwick 2007). However, direct observation of the gas turbulence from line broadening have not been detected because of the low level of turbulence (e.g., Dartois et al. 2003; Piétu et al. 2007; Hughes et al. 2011; Guilloteau et al. 2012; Flaherty et al. 2015, 2017; Teague et al. 2016, 2018), except for one tentative detection on DM Tau (Flaherty et al. 2020).

We focus on the dust scale height at the dust ring to constrain the physical state of protoplanetary disks. Pinte et al. (2016) estimated the dust scale height of HL Tau from the ring-gap contrast along the minor axis. The dust scale height depends on the ratio of the dimensionless parameter of the gas turbulence α (Shakura & Sunyaev 1973) to the dust-to-gas-coupling parameter called the Stokes number St = π ρmat adust/2Σgas, where ρmat is the dust material density, adust is the dust radius, and Σgas is the gas surface density (Epstein 1924; Adachi et al. 1976; Nakagawa et al. 1981, 1986; Garaud et al. 2004; Dullemond & Dominik 2005; Fromang & Papaloizou 2006). Therefore, we can estimate α/St by estimating the dust scale height.

In this paper, we estimate the dust scale height of HD 163296 from an ALMA dust continuum observation. The distance to this object is 101.5 pc (Gaia Collaboration et al. 2018). This object has two clear rings at 68 au and 100 au (Isella et al. 2016; Andrews et al. 2018). There is indirect evidence of a planet at 260 au based on the kinematic signature (Pinte et al. 2018) and at 78, 140, and 237 au based on the meridional flow (Teague et al. 2019). Flaherty et al. (2015, 2017) placed an upper limit on the gas turbulence of α ≤ 3 × 10−3. Guidi et al. (2016) found a trend that the dust size decreases toward the outer region based on the spectral index at millimeter wavelengths. Dullemond et al. (2018) constrained α/St based on the dust ring width and the upper limit of the gas ring width. Rosotti et al. (2020) expanded on the results of Dullemond et al. (2018) and estimated α/St by the deviation of the gas rotating velocity from the Keplerian rotation instead of from the upper limit of the gas ring width. Ohashi & Kataoka (2019) estimated the dust scale height, the dust size, and the turbulence at the gaps by millimeter-wave polarization. In this paper, we estimate the dust scale height, which becomes possible owing to the high spatial resolution by the campaign called Disk Substructures at High Angular Resolution Project (DSHARP; Andrews et al. 2018).

The structure of this paper is as follows. In Section 2 we consider a simple dust ring model to derive the general expression of the intensity as a function of the scale height. We develop a method for constraining the dust scale height and consider the conditions under which the dust scale height can be determined from observations. In Section 3 we determine the dust scale height by applying the method to the image of HD 163296. In Section 4 we discuss physical quantities such as the turbulence and the dust size, and also the ring formation mechanism based on the dust scale height.

2. Analytical Model

In this paper, we aim at determining the dust scale height of dust rings in an inclined protoplanetary disk from an observed radio continuum image. In this section, we assume a protoplanetary disk hosting an axisymmetric ring and analytically calculate the intensity of the disk. We represent the dependence of the intensity on the dust scale height if the line of sight is inclined to the disk. In Section 2.1 we describe the optical depth of the disk. In Section 2.2 we calculate the intensity at the peak positions along the major and minor axes. In Section 2.3 we discuss the conditions that allow us to determine the dust scale height.

2.1. Optical Depth of an Inclined Disk

We consider a disk hosting an axisymmetric dust ring whose radial distribution of the dust surface density and the vertical distribution of the dust spatial density are described as Gaussian functions. We take a cylindrical coordinate system as r = 0 at the central star, ϕ = 0 along the major axis, and z = 0 at the midplane. The surface density of the ring at the peak (r = r0) is Σ0, the ring width is σr , and the scale height is σz . The dust spatial density of this object is expressed as

Equation (1)

We assume the disk inclination angle to be i, and calculate the optical depth of this object along the line of sight. Figure 1 illustrates the disk geometry. In the calculation of the optical depth, we ignore the curvature of the ring. This approximation is appropriate if the ring radius is sufficiently larger than the ring width. The optical depth at r = r0 + Δr and ϕ is expressed as

Equation (2)

where κ is the dust opacity.

Figure 1.

Figure 1. Principle showing that the optical depth is different on the major and minor axes when we observed it at an oblique angle.

Standard image High-resolution image

The optical depth depends on the azimuthal angle, as shown in Equation (2). We use the word "ridge" to describe the peak position in the radial direction (r = r0) at each azimuthal angle. In particular, the optical depth along the ridge of the ring is the maximum on the major axis and the minimum on the minor axis.

2.2. Intensity on the Major and Minor Axes

We calculate the azimuthal intensity variation of both optically thin and thick disks. Here, we ignore the effects of beam smearing in interferometric observations, which we discuss in Section 2.3.

In the following discussion, we assume that the ring is isothermal in both vertical and radial directions. In reality, there are temperature gradients in both the radial and vertical directions. We assume that the dust ring width is small, so the radial temperature variation can be ignored. In the vertical direction, the temperature is higher in the upper layers that are irradiated by the central star (Chiang & Goldreich 1997; D'Alessio et al. 1998; Dartois et al. 2003; D'Alessio et al. 2005; Rosenfeld et al. 2013). For simplicity, we ignore the vertical temperature gradient. We discuss the effect of the vertical temperature gradient in Appendix A.

For optically thick disks, the intensity can be approximated as

Equation (3)

where T is the temperature of the dust ring, and Bν (T) is the Planck function. In other words, there is no variation in intensity if the ring is optically thick.

For optically thin disks, the intensity can be approximated as

Equation (4)

On the major axis(ϕ = 0°), the intensity on the ridge is

Equation (5)

and on the minor axis (ϕ = 90°), the intensity is

Equation (6)

The ratio of the intensity on the ridge on the major axis to that on the minor axis is

Equation (7)

Thus, there is a difference in intensity between the major and minor axes.

For the optically thin case, we further discuss the difference of the intensity between the major and minor axes. As shown in Equation (7), the projected dust scale height, ${\sigma }_{z}\sin i$, relative to the projected ring width, ${\sigma }_{r}\cos i$, is observed as the difference in the intensity between the major and minor axes. If ${\sigma }_{r}\cos i\gg {\sigma }_{z}\sin i$, there is no difference in the intensity between the major and minor axes. On the other hand, if the projected ring width, ${\sigma }_{r}\cos i$, is not much wider than the projected dust scale height, ${\sigma }_{z}\sin i$, the difference in the intensity between the major and minor axes depends on the dust scale height.

Figure 2 shows images of radiative transfer simulations of protoplanetary disk models composed of a dust ring with different optical and geometric thicknesses. We assume that all rings have the same width (σ = 5 au) and distance from the center (r0 = 100 au). We assume that the rings are isothermal, and we set the temperature such that the brightness temperature at the peak position on the major axis is 10 K. In the left panel, the optical depth at the ridge of the ring is κΣ0 = 5, and the temperature is 15.0 K. In the middle and right panels, the optical depth at the ridge of the ring is κΣ0 = 0.01 and the temperature is 717.8K. The scale height in the middle panel is 5/16 au, and that in the left and right panels is 5 au. Figure 3 plots the intensity along the ridges of the simulated images shown in Figure 2 and Equation (12) from the analytical calculations. We set the inclination to be 45. We can see that for the optically thick (the left panel) or geometrically thin (the middle panel) cases, the intensity does not depend on the azimuthal direction, but for the optically thin and geometrically thick case (the right panel), the intensity on the minor axis is smaller than that on the major axis. We confirm that Equation (12) describes the simulation results well. From the above, we conclude that the intensity along the ridges depends on the azimuthal angle if the ring is optically thin, the dust scale height is not too wide compared to the dust ring width, and the inclination is not too small.

Figure 2.

Figure 2. Radiative transfer images of an optically thick and geometrically thick ring (left), an optically thin and geometrically thin ring (middle), and an optically thin and geometrically thick ring (right). We assume that all of the rings are isothermal, and we set the temperature such that the brightness temperature at the peak position on the major axis is 10 K. In the left panel, the optical depth at the ridge of the ring is κΣ0 = 5, and the temperature is 15.0 K. In the middle and right panels, the optical depth at the ridge of the ring is κΣ0 = 0.01 and the temperature is 717.8K. The scale height of the middle panel is 5/16 au, and that of the left and right panels is 5 au.

Standard image High-resolution image
Figure 3.

Figure 3. Intensity along the ridges in Figure 2. The blue, green, and orange lines represent the results of the simulation, and the pink line represents the analytical solution for the optically thin and geometrically thick case.

Standard image High-resolution image
Figure 4.

Figure 4. Image of dust continuum of HD 163296 with ALMA Band 6 (λ = 1.25 mm). We use the brightness temperature, assuming the Rayleigh–Jeans law as a representation of intensity to preserve its linearity with the observed intensity.

Standard image High-resolution image

Recent studies have pointed out that optically thick emission can be fainter than the Planck function if the albedo is high (Miyake & Nakagawa 1993; Liu 2019; Zhu et al. 2019; Ueda et al. 2020; Lin et al. 2020). If the scattering works, one can imagine that the azimuthal variation may disappear due to the scattering-induced intensity reduction. This does not change our conclusion if the ring is optically thin (τabs ≪ 1) because the scattering effects are not significant. Moreover, this does not change if the ring is extremely optically thick (τabs ≫ 1) because we do not expect any azimuthal variation even if we do not consider the scattering effects. We should be careful if the ring is marginally optically thick (τabs ∼ 1). In this case, if we do not consider the effects of scattering, we expect an azimuthal intensity variation to some extent. However, if scattering plays a role the intensity on the major axis decreases more than that on the minor axis. As a result, the azimuthal variation is suppressed compared with the nonscattering case, and it would be observed as if it were extremely optically thick. Therefore, the scattering effects may diminish the azimuthal variation of marginally optically thick rings. Note that the scattering may not significantly contribute to the emission in the case of HD 163296 because τext = τabs + τscat < 1 from the observation of CO backside emission (Isella et al. 2018).

2.3. Conditions for Estimating Dust Scale Heights from Observations

We discuss the conditions in which we can constrain the dust scale heights, including the effects of the beam smearing of the interferometric observations. We approximate the beam smearing as a Gaussian convolution of an ellipsoidal shape. Based on the previous section, we only consider the optically thin case because we are not able to constrain the dust scale height if the dust ring is optically thick. From Equation (2), the intensity at the location (r + Δr, ϕ) can be expressed as

Equation (8)

We convolve this equation with a Gaussian beam. We deproject the observed Gaussian beam for the inclination and denote the standard deviation of the deprojected beam along the ϕ direction as σbeam(ϕ). Here we ignore the curvature of the ring and convolve the radial profile and the beam smearing. The convolved intensity is

Equation (9)

Particularly, the intensities along the major axis and the minor axis are expressed as

Equation (10)

Equation (11)

Along the ridge of the ring, the intensity is expressed as

Equation (12)

From Equation (10), the intensity along the major axis is independent of the dust scale height. Therefore, we can construct a disk model that is independent of the dust scale height from the observed intensity along the major axis. On the other hand, from Equations (11) and (12), the intensity along the ridge except for the major axis or along the minor axis depends on the dust scale height. Therefore, we can constrain the dust scale height by the consistency of the observed intensity along the ridge or the minor axis.

We discuss the conditions for constraining the dust scale height from the observed images. The dust scale height contributes to the intensity as the term ${\sigma }_{z}^{2}{\tan }^{2}i$ in the denominator of Equations (11) and (12). If ${\sigma }_{z}^{2}{\tan }^{2}i$ is not too small compared to the other terms (${\sigma }_{r}^{2}$, ${\sigma }_{\mathrm{beam}}^{2}$), the effect of the dust scale height appears in the observed intensity. Then, we can constrain the dust scale height from observations. We further discuss the effect of the beam smearing in Appendix B.

To summarize the discussion above, we can constrain the dust scale height in the following steps. First, we construct a disk surface density model from the observed intensity along the major axis. Then, we search for the dust scale height that reproduces the observed intensity along the ridge or the minor axis. We require the following four conditions to constrain dust scale heights: 1. The dust ring is optically thin, 2. the dust ring width is not much wider than the dust scale height, 3. the inclination is not too small, and 4. the dust ring is spatially resolved.

3. Application to HD 163296

We apply the method above to the protoplanetary disk around HD 163296 to constrain the dust scale height by comparing the observation and radiative transfer simulations. In Section 3.1 we describe the observational data used in this work. In Section 3.2 we describe the disk model for the radiative transfer simulation. In Section 3.3 we compare the azimuthal variation of the intensity along the ridge as the first comparison method. In Section 3.4 we compare the radial intensity profile along the major and minor axes as the second comparison method.

3.1. Observation

In this study, we use the fits data of the image of the HD 163296 disk taken in the DSHARP campaign (2016.1.00484.L), which is one of the ALMA large programs (Andrews et al. 2018). We show the image in Figure 4. This image was taken with ALMA Band 6, which corresponds to the wavelengths of λ = 1.25 mm. This image was made by combining the data of project 2013.1.00366.S13 (Flaherty et al. 2015) and project 2013.1.00601.S4 (Isella et al. 2016) to improve the sensitivity and uv-coverage on shorter baselines. The details of this image are shown in Andrews et al. (2018) and Isella et al. (2018). We assume that the distance to this object is 101.5 pc from Gaia DR2 parallax measurements (Gaia Collaboration et al. 2018). We assume that the central star mass Mstar = 2.02Msun is a dynamical mass based on the line observation from Teague et al. (2019).

This object satisfies the requirements for constraining the dust scale height described in Section 2.3. The observation's beam size is 0farcs020 × 0farcs016 with a standard deviation that corresponds to 2.1 × 1.6 au, and the beam position angle is 81fdg7 (Isella et al. 2018). This source has narrow rings; the width is 4.0 au for the inner ring and 3.9 au for the outer ring with a standard deviation, which we discuss in Section 3.2. The disk inclination is 46fdg7 ± 0fdg1, and the disk position angle is 133fdg1 ± 0fdg1 (Isella et al. 2018). The gas scale heights of this source are 3.8 au in the inner ring and 6.7 au in the outer ring, using the temperature model discussed in Section 3.2. As described above, this object is sufficiently inclined, and the ring width and beam are not too large compared to the gas scale height, so we think that we can constrain the dust scale height of this source.

3.2. Setup of the Simulation

We describe our simulation setup to compare with the observation. We perform radiative transfer simulations with RADMC-3D (Dullemond et al. 2012). In our simulation, we use spherical coordinates (r, θ, ϕ), and we assume axial symmetry in ϕ direction and plane symmetry with respect to the midplane. The radial grid is linearly spaced between rin = 10 au and rout = 200 au with 512 grids. The θ grid is linearly spaced with 512 grids in the range 0 ≤ θ ≤ 0.2, where θ = 0 is the midplane, and θ = π/2 is the north pole.

We adopt the following temperature profile with the smooth power-law distribution derived by Dullemond et al. (2020),

Equation (13)

Dullemond et al. (2020) discussed the midplane temperature based on the CO line emission. We assume that the disk is isothermal in the vertical direction. In reality, the disk surface is hotter than the midplane, but the vertical distribution of the gas is approximated well by hydrodynamic equilibrium at midplane temperature, so using the midplane temperature is a good approximation when we consider the vertical distribution of the gas. We discuss the effect of the vertical temperature gradient on the intensity in Appendix A.

We use the dust opacity model of Birnstiel et al. (2018) with κabs = 0.484 cm2 g−1 at a wavelength of λ = 1.25 mm. Here, we consider only the absorption of the dust and do not include dust scattering (see Miyake & Nakagawa 1993; Kataoka et al. 2014; Liu 2019; Zhu et al. 2019; Ueda et al. 2020; Lin et al. 2020). We need the dust surface density and dust opacity for the input parameters of RADMC-3D. From the observed intensity and temperature distribution model of Equation (13), we can derive the optical depth but cannot solve the degeneracy of the dust surface density and dust opacity, so we need a dust opacity model to derive the dust surface density. However, output images depend on the optical depth regardless of the combination of the dust opacity and surface density. Therefore, our result does not depend on the dust opacity model.

We make a surface density profile based on the observational radial intensity profile along the major axis. This is possible because the intensity along the major axis does not depend on the dust scale height, as shown in Equation (10). We use the radial intensity profile along the major axis in the northwest direction to avoid the additional crescent-shaped structure reported in the southeast direction (Isella et al. 2018). We perform numerical simulations to search for parameters that reproduce the observational intensity along the major axis. We assume that the dust surface density model is composed of an inner disk and two Gaussian rings as

Equation (14)

We use the parameters Σdisk = 20.66 g cm−2, rdisk = 25.0 au, Σring1 = 1.26 g cm−2, r0,ring1 = 67.9 au, wring1 = 4.0 au, Σring2 = 0.58 g cm−2, r0,ring2 = 100.5 au, and wring2 = 3.9 au. In Section 3.4 we discuss the consistency of this model and observation. We summarize the parameters related to the rings on which we focus in this study in Table 1.

Table 1. Parameters of the Dust Rings of HD 163296

  r0 Σdust wdust hgas
Ring(au)(g cm−2)(au)(au)
(1)(2)(3)(4)(5)
inner ring67.91.264.03.8
outer ring100.50.583.96.7

Note. Column (2): Distance to the central star. Column (3): The dust surface density at the peak of the rings. Column (4): The width of the dust rings. Column (5): Gas scale height based on midplane temperature and mass of the central star.

Download table as:  ASCIITypeset image

The absorption optical depth (${\tau }_{\mathrm{abs}}={\kappa }_{\mathrm{abs}}{{\rm{\Sigma }}}_{d}/\cos i$) of this model at the rings is 0.89 for the inner ring and 0.41 for the outer ring. Isella et al. (2018) estimated the extinction optical depth, τext, of this source from the extinction of CO backside emission at the rings, and constrained that τext = 0.64 ± 0.05 for the inner ring and τext = 0.74 ± 0.05 for the outer ring. Since κext = κabs + κscat, τext must be higher than τabs. However, τabs in this study and the constraint of τext in Isella et al. (2018) do not satisfy this relation for the inner ring. A possible explanation for this inconsistency is that the dust surface density is overestimated because the temperature model used in this study underestimates the temperature, or τext in Isella et al. (2018) is underestimated due to beam smearing or contamination of frontside emission.

Now we are ready to determine the dust scale height. To estimate the scale height that reproduces the observation, we change the dust settling parameter fset. We define the dust settling parameter such that fset = hgas/hdust, where hdust is dust scale height and hgas is the gas scale height. The gas pressure scale height is hgas = cs kep from the hydrostatic equilibrium, where cs is the isothermal sound speed and Ωkep is the Keplerian orbital velocity. We run multiple simulations in the range of 1/2 ≤ fset ≤ 16 with every 21/16 steps, which corresponds to 81 models. We convolve the simulated images with the beam and compare them with the observed image.

3.3. Comparison of the Intensity: The Azimuthal Variation along the Ridge

The analytical expression of the azimuthal variation described as Equation (12) shows that the intensity along the ridge depends on the dust scale height. We constrain the dust scale height by comparing the azimuthal intensity variation along the ridge of the observation with that of the simulations.

To demonstrate the azimuthal intensity variation, we show in Figure 5 images of the observation and simulations in the case of fset = 1 (no settling) and 8 (settled). We use a different color range for each row to emphasize the azimuthal variations of the inner and outer rings. Figure 6 shows the azimuthal variation along the ridges of the observation and simulations in the case of fset = 1, 2, and 8. As shown in the simulation images of fset = 1, where the dust is flared, the intensity along the ridge is higher on the major axis (ϕ = 0°, 180°) than on the minor axis (ϕ = 90°, 270°). In contrast, as shown in the simulation images of fset = 8, where the dust is settled, the intensity along the ridge does not depend on the azimuthal angle. The observational images show that the intensity along the ridge of the inner ring is lower on the minor axis (ϕ = 90°, 270°) than on the major axis (ϕ = 0°, 180°). In contrast, the intensity along the ridge of the outer ring is uniform for the azimuthal direction. This means that the dust scale height is large for the inner ring and small for the outer ring.

Figure 5.

Figure 5. Comparison of images of HD 163296. The three images in the left column represent observations, and the three images in the middle and right columns represent simulations with fset = 1 (no settling) and fset = 8, respectively. Each line shows the images with different color ranges. The middle line shows the images that emphasize the azimuthal intensity variation of the inner ring, and the bottom line shows that of the outer ring.

Standard image High-resolution image
Figure 6.

Figure 6. Azimuthal intensity variation along the ridge. The red dots show the observations, and the solid lines represent the simulations. The black, blue, and cyan line indicate fset = 1 (no settling), 2 and 8, respectively. The shaded regions represent the areas with non-axisymmetric features such as a crescent. We plot the data at every FWHM of the beam (Huang et al. 2018).

Standard image High-resolution image

We constrain fset using the χ2. We run simulations varying fset from 1/2 to 16 by a factor of 21/16, and calculate χ2. To estimate the parameter fset, we calculate the χ2 between the observation and the simulations along the ridge. To exclude the effect of the crescent-shaped substructure in the southeast direction, we excluded the region between − 45° and 45 in our calculations of the χ2. Figure 7 shows the χ2 at each fset. The best-fit parameters are fset = 1.1 for the inner ring and fset = 16 for the outer ring. We search the range where ${\chi }^{2}\lt {\chi }_{{\rm{\min }}}^{2}+{3}^{2}$. We obtain ${f}_{{\rm{set}}}={1.1}_{-0.1}^{+0.1}$ for the inner ring and fset > 9.5 for the outer ring with 3σ uncertainty. With a 3σ uncertainty, the inner ring is more flared than the outer ring. Figure 7 also shows that our method has high sensitivity where ${f}_{{\rm{set}}}$ is small while it has low sensitivity where ${f}_{{\rm{set}}}$ is large. Note that ${f}_{{\rm{set}}}$ cannot be smaller than unity because the dust scale height cannot be larger than the gas scale height, while our results satisfy this condition.

Figure 7.

Figure 7.  χ2 of the intensity between the observation and simulations along the ridge for each fset. We use data at every FWHM. We exclude data for the region with a crescent structure (shaded area in Figure 6).

Standard image High-resolution image

3.4. Consistency with the Radial Profiles

We discuss whether the derived dust scale height from the azimuthal intensity variation is consistent with the observed radial intensity profile along the minor axis. From Equations (10) and (11), the intensity along the major axis does not depend on the dust scale height, while that along the minor axis depends on the dust scale height. Along the minor axis, the higher the dust scale height, the wider the observed dust ring width and the lower the observed intensity at the peak. We plot the intensity of both simulations and the observation along the major and minor axes to check the consistency of fset, which we estimate in Section 3.3.

Figure 8 shows the radial intensity profiles of the observation along the major and minor axes. Around the inner ring at 68 au, the intensity along the major axis is higher than that along the minor axis. On the other hand, around the outer ring at 100 au, the intensity along the major and minor axes is almost the same.

Figure 8.

Figure 8. Radial intensity profiles of the observation averaged over 6° width from the major and minor axes. The blue and purple lines show the intensity along the major axis, and the orange and red lines represent that along the minor axis.

Standard image High-resolution image

Figure 9 shows the radial intensity profiles along the major (top) and minor (bottom) axes of the observation and simulations in case of fset = 1 and 8. We can see that the simulations reproduces the observation well along the major axis from the upper panel of Figure 9. The bottom panel of Figure 9 shows that the radial profile along the minor axis around the inner ring is well reproduced by fset = 1, while that around the outer ring is well reproduced by fset = 8. This result is consistent with the result in Section 3.3 that the inner ring is flared and the outer ring is settled.

Figure 9.

Figure 9. Radial intensity profiles along the major axis (top) and minor axis (bottom). The black and gray lines are the intensity of the simulations, and the other lines are the intensity of the observations. The intensity along the major axis does not depend on the dust scale height, so the two simulation lines overlap.

Standard image High-resolution image

Figure 10 shows the radial intensity profile of the best model along the minor axis. Based on the result in Section 3.3, we set fset = 1.1 for the inner ring and fset = 16 for the outer ring. This result reproduces the onservation well for the entire region.

Figure 10.

Figure 10. Radial intensity profiles along the minor axis. The black line represents the intensity of the best model (fset = 1.1 for the inner ring and fset = 16 for the outer ring), and the blue and cyan lines are the intensity of the observation.

Standard image High-resolution image

4. Discussion

4.1. The Dust Scale Height for the Entire Region

We discuss the dust scale height for the entire region of the protoplanetary disk. Figure 11 shows a schematic illustration of the dust scale height of HD 163296. Ohashi & Kataoka (2019) determine the dust scale height at the gaps by using the polarized intensity of this object. They concluded that the dust scale height is less than one-third of the gas scale height at the gap at 48 au, and the dust scale height is two-thirds of the gas scale height at the gap at 86 au. Taken together with the results of this study, we can draw the entire picture of the radial distribution of the dust scale height except for the central disk. From the inner side to the outer side, the scale height is unknown at the central disk, thin at the 48 au gap, thick at the 68 au ring and the 87 au gap, and thin at the 100 au ring.

Figure 11.

Figure 11. Schematic view of the dust scale height for the entire region. The scale height in the gaps is taken from Ohashi & Kataoka (2019), and the scale height in the rings is from our study. The dust scale height of the central disk remains unknown.

Standard image High-resolution image

We discuss whether the radial variation of the dust scale height can be explained by a simple dead-zone model (Gammie 1996; Okuzumi et al. 2016; Ueda et al. 2019). The dead zone exists in the inner part of the protoplanetary disk because the low-ionization ratio in the inner part of the disk suppresses the magnetorotational instability. The dust is settled in the dead zone due to weak gas turbulence. Therefore, the dust scale height is smaller in the dead zone and larger outside the dead zone. As shown in Figure 11, however, the dust scale height of HD 163296 varies in a complex way in the radial direction and cannot be explained by a simple disk model with a dead zone. This suggests that other factors cause the change in dust scale height, e.g., the radial change of the dust size or the turbulence.

We compare our dust scale height model with an infrared polarization observation. Muro-Arena et al. (2018) made a 3D disk model of HD 163296 using millimeter dust continuum observation using ALMA and infrared polarized scattered light using SPHERE. They reported that the polarized infrared light is observed only at the inner ring at 68 au and is not observed at the outer ring at 100 au. This result can be interpreted to mean that the outer ring is in the shadow of the inner ring, which is consistent with our results of the dust scale height.

4.2.  α/St

We estimate α/St from the dust scale height. By assuming a balance between vertical settling and turbulent diffusion, the dust scale height is written as (e.g., Dubrulle et al. 1995; Youdin & Lithwick 2007)

Equation (15)

where h is the scale height, and the subscripts g and d indicate gas and dust, respectively. This equation can be approximated as

Equation (16)

if St ≪ 1. Using the fset constrained in Section 3.3, we derive α/St > 2.4 for the inner ring and α/St < $1.1\times {10}^{-2}$ for the outer ring with 3σ uncertainties. We note that Equation (16) may not be appropriate if ${h}_{d}\approx {h}_{g}$.

We discuss the consistency with measurements in other studies. Dullemond et al. (2018) and Rosotti et al. (2020) estimated α/St by the radial diffusion of dust due to the gas turbulence. Dullemond et al. (2018) assumed ring formation by dust traps at gas bumps and derived the following relationship between dust ring width wd and gas ring width wg :

Equation (17)

Dullemond et al. (2018) estimated the dust ring width from the dust continuum emission, and Rosotti et al. (2020) estimated the gas ring width from the deviation of the gas rotation velocity. Their results are α/St = 0.23 for the inner ring and α/St = 0.04 for the outer ring. We note that Rosotti et al. (2020) and our results show the same tendency that the inner ring is more diffuse than the outer ring. However, their $\alpha /{\rm{St}}$ is somewhat different from that of us. There are two possible reasons for this disagreement. One is that the turbulence is not isotropic, i.e., the turbulence is different in radial and vertical directions. The other is that the ring formation mechanism is different from that assumed by Rosotti et al. (2020). Rosotti et al. (2020) assume that ring formation is due to gas bumps. In other scenarios, however, the dust ring width and gas ring width do not necessarily satisfy the relationship of Equation (17). We discuss other ring formation mechanisms in Section 4.5.

4.3. Dust Size

In this section, we constrain the dust size based on the α/St. In Section 4.2 we obtained α/St based on fset. The Stokes number is expressed as

Equation (18)

where ρmat is the dust material density, adust is the dust radius, and Σgas is the gas surface density. Therefore, we can constrain the dust size from α/St obtained in Section 4.2 and α, Σgas, and ρmat. We assume that the dust is compact and icy (ρmat = 1.0 g cm−3), and the gas surface density is Σgas = 21.1 g cm−2 for the inner ring and Σgas = 10.0 g cm−2 for the outer ring, following Rab et al. (2020) based on molecular line observations. Flaherty et al. (2017) placed an upper limit of α < 3 × 10−3, so we calculate the dust radius for the typical three turbulences as α = 3 × 10−3, 1 × 10−3, and 1 × 10−4. We show the results in Table 2. If we assume α = 1 × 10−3, we obtain adust < 5.6 × 10−3 cm for the inner ring and adust > 5.7 × 10−1 cm for the outer ring.

Table 2. Estimation of the Dust Size at Typical Gas Turbulence

Ring fset α/St Σgas α adust
   (g cm−2 ) (cm)
(1)(2)(3)(4)(5)(6)
    3 × 10−3 <1.7 × 10−2
inner ring<${1.1}_{-0.1}^{+0.1}$ >2.421.01 × 10−3 <5.6 × 10−3
    1 × 10−4 <5.6 × 10−4
    3 × 10−3 >1.7
outer ring>9.5<1.1 × 10−2 10.01 × 10−3 >5.7 × 10−1
    1 × 10−4 >5.7 × 10−2

Note. Column (2): Settling parameters that we constrain in Section 3.3. Column (3): α/St that we constrain in Section 4.2. Column (4): Gas surface density at the peak position of the ring (Rab et al. 2020). Column (5): Gas turbulence. We assume three typical parameters based on the upper limit from Flaherty et al. (2017). Columns (6): Dust size calculated from the left parameters.

Download table as:  ASCIITypeset image

Since Flaherty et al. (2017) placed an upper limit of α < 3 × 10−3, we can constrain the dust size of the inner ring without any assumption of turbulence. As shown in Table 2, the dust size in the inner ring is adust < 1.7 × 10−2cm. The result rules out models in which the midplane dust consists of centimeter-sized pebbles. Note that Flaherty et al. (2017) assumed that α is constant throughout the disk, but this may not be the case.

If the turbulence is the same in the two rings, we find that the dust size is larger in the outer ring than in the inner ring. Guidi et al. (2016) found a trend of the dust size from the spectral index. They concluded that the dust size decreases toward the outside from the increasing trend of the spectral index, while the spatial resolution is not high enough to resolve the rings. Dent et al. (2019) obtained the specrtal index, αSED, between Band 6 and Band 7 (0.87 mm) from high-resolution observations that spatially resolve the ring. Dent et al. (2019) obtained that αSED in the rings is smaller than that in the gaps, and αSED in the inner ring (αSED = 2.2) is larger than that in the outer ring (αSED = 1.8). Since the systematic uncertainty is large, αSED for the total disk flux differs among Pinilla et al. (2014; 2.73 ± 0.44), Guidi et al. (2016) and Notsu et al. (2019; 2.7), and Dent et al. (2019; 2.1 ± 0.3). The value of αSED itself may not be accurate, but we can consider that the radial trend of αSED is correct. If we assume that this disk is optically thin, the dust is smaller in the inner ring than in the outer ring. This result is consistent with our result assuming the same turbulence between the two rings. We do not rule out the possibility that the gas turbulence is different between the inner and outer rings. We note that if the outer ring is optically thick, αSED becomes smaller in the outer ring, and the azimuthal variation in intensity is small.

Ohashi & Kataoka (2019) estimated the maximum dust size at the dust gap to be 140 μm based on the polarization fraction of ALMA Band 7 (0.87 mm) dust continuum. This result is smaller than the dust size of the outer ring shown in Table 2. The dust size estimated from the polarization observation is just the dust size in the gap, and we cannot directly compare it with the dust size in the ring. As shown in Table 2, however, it is consistent to assume that the dust size in the inner ring is the same as that in the gaps.

4.4. Turbulence

In this section, we discuss the possible range of α and St. Since our method cannot solve for the degeneracy of α and St from α/St, we estimate α under the following two assumptions: one assumption is that the dust size is limited by the collisional fragmentation due to the turbulence, and the other assumption is that the dust size is fixed to 1 mm.

First, we consider the case where the dust size is limited by the collisional fragmentation of dust grains induced by gas turbulence. In this case, α can be written as (Birnstiel et al. 2012; Rosotti et al. 2020)

Equation (19)

where vfrag is the fragmentation velocity and cs is the sound speed isothermal. We obtain αfrag > 3.0 × 10−2 for the inner ring and αfrag < 2.1 × 10−3 for the outer ring when we assume the fragmentation velocity to be vfrag = 10 m/s. The dust size may not be determined by the fragmentation limit by the turbulence, but if the turbulence is higher than αfrag, the dust is destroyed and becomes smaller. Therefore, αfrag is the upper limit of the turbulence.

Next, we consider the turbulence when the dust size is 1 mm. As we discuss in Section 4.3, the dust size may not be the same for the two rings, but as an example, we calculate the turbulence α1 mm when the dust size adust = 1 mm for the two rings. The St is determined as Equation (18), and we assume that the gas surface density and dust material density are the same as those in Section 4.3. Under these assumptions, we obtain α1 mm > 1.8 × 10−2 for the inner ring and α1 mm < 1.8 × 10−4 for the outer ring.

We summarize the results in Table 3 and Figure 12. The constraint of α in this study at the inner ring is larger than the observational upper limit of α in Flaherty et al. (2017). There are three possible explanations for this inconsistency. First, there may be a mechanism that keeps the dust size small in the inner ring. For example, the fragmentation velocity may be lower than the assumption in this study due to dust sintering (Okuzumi et al. 2016). Second, the turbulence may vary in the radial direction. Flaherty et al. (2017) constrained α throughout the disk from gas emission line observations, and this value corresponds to the intensity-weighted average of α. Since the gas disk is larger than the dust ring, low turbulence in the outer side results in a small α throughout the disk. Third, the α of the dust diffusion we estimate may be different from the α of the gas motion measured by Flaherty et al. (2017). Pinilla et al. (2021) discuss the case where α for gas evolution, turbulent velocity, vertical diffusion, and radial diffusion takes different values.

Figure 12.

Figure 12. Possible ranges of α and St. The blue and orange areas represent the possible ranges of α and St for the inner and outer rings, respectively. The dashed red line represents the fragmentation limit. The dashed brown line represents the upper limit of α from the observation of line broadening (Flaherty et al. 2017).

Standard image High-resolution image

Table 3. Turbulence α under Certain Assumptions

Ring α/St αfrag α1 mm
(1)(2)(3)(4)
inner ring>2.4>3.0 × 10−2 >1.8 × 10−2
outer ring<1.1 × 10−2 <2.1 × 10−3 <1.8 × 10−4

Note. (3) Turbulence α where the dust size is limited by the collisional fragmentation due to the turbulence. (4) Turbulence α when the dust size is 1 mm.

Download table as:  ASCIITypeset image

Liu et al. (2018) constrained α from hydrodynamic simulations assuming gap opening by planets. They showed that if α is small ( ≤ 10−4) in the inner disk and large ( ≥ 7.5 × 10−3) in the outer disk, they can reproduce both dust and gas observations. However, our study shows that α/St is larger at the inner ring than at the outer ring, which is different from Liu et al. (2018). This disagreement indicates that it is difficult to reproduce the gas and dust profiles with the dust trap alone. The ring formation mechanism is discussed in Section 4.5.

4.5. Ring Formation Mechanism

In this section, we discuss the ring formation process based on the dust scale height that we estimate in this study. Previously, several ring formation mechanisms have been proposed, and we consider the following four mechanisms: 1. Dust accumulation by gas gaps induced by planets, 2. enhanced fragmentation of dust grains by sintering at snowlines, 3. dust accumulation at the outer edge of the dead zone, and 4. secular gravitational instability (SGI). This study shows that the settling conditions of the two rings are different. Therefore, we consider ring formation mechanisms of each ring independently.

A possible mechanism is dust accumulation by the pressure gaps induced by planets (Lin & Papaloizou 1979; Goldreich & Tremaine 1980; Duffell & MacFadyen 2013; Fung et al. 2014; Kanagawa et al. 2015). Zhu et al. (2012) show that large dust accumulates on the outside of the pressure gap, while small dust couples with gas and flows into the inside of the pressure gap. This process is called dust filtration. Since the rings formed by this process consist of large dust, the dust is settled in the midplane. Compared to the results of our study, the outer ring is consistent with this process because the dust is settled, while the inner ring is inconsistent with this process because the dust is flared.

Another possible mechanism is dust sintering (Okuzumi et al. 2016; Zhang et al. 2015). The sintering theory predicts that the rings consist of small dust fragments. The inner ring may be explained by dust sintering because the dust is flared, but the outer ring cannot be explained because the dust is settled, which suggests large dust grains.

Another possible mechanism is dust accumulation at the outer edge of the dead zone (Flock et al. 2015; Pinilla et al. 2016; Mori & Okuzumi 2016; Ueda et al. 2019). The dead zone is the inner region of the disk where the MRI is stabilized and the gas turbulence is weak. However, since the dust in the inner ring is not settled, the turbulence is not weak, and the dead zone is not the formation mechanism of the ring. The outer ring is far from the central star when we consider the dead zone. We do not rule out the possibility that the central disk is due to the dead zone.

The last possible mechanism we consider is SGI (Ward 2000; Youdin 2011; Michikoshi et al. 2012; Takahashi & Inutsuka 2014; Tominaga et al. 2019). The SGI is stabilized by the turbulence, so the turbulence must be weak for the instability to develop. Therefore, the dust is settled in the dust ring that is formed by the SGI. From the perspective of the dust scale height, the outer ring is consistent, and the inner ring is inconsistent.

To summarize the above, from the perspective of the dust scale height, the formation mechanism of the inner ring is consistent with dust sintering. In contrast, the formation mechanism of the outer ring is consistent with dust accumulation by gas gaps induced by planets and secular gravitational instability. We do not rule out other ring-forming mechanisms, such as dust replenishment from a planet around the substructure.

4.6. Caveats

We assume the temperature to be isothermal in the vertical direction and a simple power law in the radial direction, but this may be different from the actual temperature.

In this study, we focus on only one object, HD 163296. This method can be applied to other objects to estimate the dust scale height. By using high spatial resolution observations, it is possible to limit the dust scale height of other objects.

5. Conclusion

We developed a method for estimating the dust scale heights of dust rings of protoplanetary disks from dust continuum images. We apply the method to the DSHARP dust continuum image of HD 163296 (Andrews et al. 2018). Based on the scale height we constrain, we discuss the physical state of the rings and the ring formation mechanisms. Our conclusions are as follows.

  • 1.  
    The intensity of the dust rings depends on the dust scale height, especially around the minor axes. Therefore, we can estimate the dust scale height by comparing the intensity along the ridges of the rings. To constrain the dust scale height, we need the following conditions: 1. The dust ring is optically thin, 2. the dust ring width is not much wider than the dust scale height, 3. the inclination is not too small, and 4. the dust ring is spatially resolved.
  • 2.  
    We constrain the dust scale height of HD 163296. The dust scale height is different between the inner ring at 68 au and the outer ring at 100 au. We obtain ${f}_{{\rm{set}}}={1.1}_{-0.1}^{+0.1}$ for the inner ring and fset > 9.5 for the outer ring with 3σ uncertainties. This indicates that the dust is flared in the inner ring and settled in the outer ring.
  • 3.  
    We discuss the dust scale height for the entire region of this object by combining this study, which constrains the dust scale height of the dust ring, and Ohashi & Kataoka (2019), who constrained that of the dust gap. The dust scale height varies in a complex way in the radial direction, and a simple dead-zone model cannot explain the variation in the dust scale height.
  • 4.  
    Based on the dust scale height, we estimate α/St. We obtain α/St > 2.4 in the inner ring, while α/St < $1.1\times {10}^{-2}$ in the outer ring. These results show that the turbulence is stronger or the dust is smaller at the inner ring than at the outer ring.
  • 5.  
    We constrain the dust size based on the derived α/St. If we assume that the turbulence is the same in the two rings, we find that the dust size is larger in the outer ring than in the inner ring. We constrain the dust size in the inner ring adust < 1.7 × 10−2 cm by using the upper limit of the gas turbulence by Flaherty et al. (2017).
  • 6.  
    We discuss the possible range of α and St. We assume a fragmentation limit and obtain that αfrag > 3.0 × 10−2 for the inner ring and αfrag < 2.1 × 10−3 for the outer ring. The αfrag in the inner ring is larger than the upper limit of α from the line broadening of α < 3 × 10−3 (Flaherty et al. 2017). We need other mechanisms to keep the dust small for consistency of our results and the constraints from the line broadening (Flaherty et al. 2017).
  • 7.  
    We discuss the dust ring formation mechanisms from the perspective of the dust scale height. The formation mechanism of the inner ring is consistent with enhanced dust fragmentation by dust sintering, while that of the outer ring is consistent with dust accumulation by a planet inducing gas gap or secular gravitational instability.

The authors thank the anonymous referee for their helpful comments. The authors thank T. Tsukagoshi, T. Ueda, and Y. Yamato for fruitful discussion and comments. This project was first started with the dust continuum data of Isella et al. (2016), which were kindly provided by A. Isella. This work was supported by JSPS KAKENHI Grant Numbers 18K13590 and 19H05088.

Appendix A: Effect of the Vertical Temperature Gradient

In this paper, we consider vertically isothermal disks. In reality, the surface layer, which is irradiated by the central star, is hotter than the midplane. We discuss how the vertical temperature structure affects the observed images. Note that we ignore the effect of accretion heating because we focus on the outer part of the disk (Chiang & Goldreich 1997; Balbus & Hawley 1998).

The dust temperature below one gas scale height is almost the same as that at the midplane (D'Alessio et al. 1998). Most of the dust is located where the height from the midplane is less than one gas scale height even if the dust is flared. Therefore, the hot upper layer hardly affects the dust continuum image, and it is reasonable to approximate that the disk is vertically isothermal.

The vertical temperature gradient can affect the observational image if the temperature varies significantly near the midplane. If the disk is optically thick, the contribution of radiation from the hot upper layer is large. Since the optical depth is larger on the major axis than on the minor axis, we observe radiation from higher layers on the major axis than the minor axis. Therefore, we expect the dust ring to be brighter on the major axis than on the minor axis. This trend of the azimuthal variation is the same as in the case of the optically thin and geometrically thick ring discussed in Section 2.2. We tested whether the above two conditions are confusing.

Even though the vertical temperature variation around the midplane is believed to be small, we demonstrate the case where the disk is optically thick and have a strong vertical temperature gradient even around the midplane. We performed additional radiative transfer simulations with the temperature model as

Equation (A1)

where Tm = 21.6 K, Ta = 77.8 K, δ = 2.16, and zq = 25.2 au. These equations are based on the temperature model in Rosenfeld et al. (2013) at 100 au. We assume that the dust is radially isothermal to focus on the effect of the vertical temperature gradient. This model leads to a temperature at one gas scale height that is almost twice as high as the midplane temperature, which may make some difference in the final images. The gas scale height is 6.6 au at r = 100 au based on the midplane temperature. We assume rings with a width, σw , of 10 au and an optical depth, κΣd , of 5.

Figure 13 shows the simulation results for fset = 1 and 2, and Figure 14 shows the intensity along the ridge. The intensity is higher than in Figure 2 if fset = 1 because the temperature at the τ = 1 surface is higher than that at the midplane. On the other hand, the intensity is hardly affected by the vertical temperature gradient if fset = 2 because the temperature on the τ = 1 surface is almost the same as that at the midplane. The intensity along the ridge is slightly brighter on the major axis than on the minor axis for fset = 1 and almost flat for fset = 2. In any case, this pattern cannot be confused with the pattern for the optically thin and geometrically thick case.

Figure 13.

Figure 13. Images that demonstrate the effect of the vertical temperature gradient. We use the temperature model of Rosenfeld et al. (2013). We assume rings with a width, σw , of 10 au and an optical depth, κΣd , of 5. The left panel shows the result for fset = 1, and the right panel shows the result for fset = 2.

Standard image High-resolution image
Figure 14.

Figure 14. Intensity along the ridges in Figure 13.

Standard image High-resolution image

Appendix B: Effect of Beam Smearing

We discuss the requirements of the spatial resolution to constrain the dust scale height. We assume that the observed image is smoothed by an ellipsoidal Gaussian function, as shown in Equation (9). The beam smearing decreases the intensity along the ridge of the ring and can cause azimuthal variations of the intensity, especially if the beam is elongated.

We convolve the model images with beams and discuss the requirements for distinguishing the difference of the dust scale height. Figure 15 shows convolved images, and Figure 16 shows the intensity of these images along the ridges. The ring models are the same as those in the middle and right panels of Figure 2. Figure 15, panels (a-1), (b-1), (c-1), and (d-1) (i.e., the four upper panels) show the geometrically thin ring models, and Figure 15, panels (a-2), (b-2), (c-2), and (d-2) (i.e., the four lower panels) show the geometrically thick ring models. Figure 15, panels (a-1), (a-2), (b-1), and (b-2) (i.e., the four left panels) show images convolved with large beams (${\sigma }_{\mathrm{beam}}=5\sqrt{2}\ \mathrm{au}$ along the major axis and σbeam = 5 au along the minor axis), and Figure 15, panels (c-1), (c-2), (d-1), and (d-2) (i.e., the four right panels) show images convolved with small beams (σbeam = 1.5 au along the major axis and ${\sigma }_{\mathrm{beam}}=1.5/\sqrt{2}\ \mathrm{au}$ along the minor axis). The major axes of the beams and the rings are aligned in Figure 15, panels (b-1), (b-2), (d-1), and (d-2), and the deprojected beams are perfect circles. The major axes of the beams and the rings are perpendicular in Figure 15, panels (a-1), (a-2), (c-1), and (c-2), and the major axes of the deprojected beams are twice as large as the minor axes.

Figure 15.

Figure 15. Images after beam smearing. Panels (a-1), (b-1), (c-1), and (d-1) (i.e., the four upper panels) are the geometrically thin ring models, and panels (a-2), (b-2), (c-2), and (d-2) (i.e., the four lower panels) are the geometrically thick ring models. Panels (a-1), (a-2), (b-1), and (b-2) (i.e., the four left panels) show images convolved with large beams (${\sigma }_{\mathrm{beam}}=5\sqrt{2}\ \mathrm{au}$ along the major axis and σbeam = 5 au along the minor axis), and panels (a-2), (b-2), (c-2), and (d-2) (i.e., the four right panels) show images convolved with small beams (σbeam = 1.5 au along the major axis and ${\sigma }_{\mathrm{beam}}=1.5/\sqrt{2}\ \mathrm{au}$ along the minor axis). The position angles of the beams and the rings are aligned in panels (b-1), (b-2), (d-1) and (d-2), and they are perpendicular in panels (a-1), (a-2), (c-1) and (c-1).

Standard image High-resolution image
Figure 16.

Figure 16. Intensity along the ridges in Figure 15.

Standard image High-resolution image

Figure 15, panel (a-1), shows that the beam smearing makes the azimuthal intensity variations even though the ring is geometrically thin. Therefore, we cannot simply conclude that the dust scale height is large from the presence of the azimuthal intensity variation if the beam size is larger than the scale height. The azimuthal intensity variation in the left panel of Figure 16 is almost the same regardless of whether the dust scale height is large or small, and we cannot constrain the dust scale height under such a beam.

Figure 15, panel (b-1), shows that the beam does not cause the azimuthal intensity variation even though the beam size is the same as that of Figure 15, panel (a-1). We can also distinguish differences in the dust scale heights from the azimuthal intensity variation in the left panel of Figure 16.

Figures 15, panels (c-1), (c-2), (d-1), and (d-2) show that the beam smearing hardly affects observational images if the beam size is small. We can clearly distinguish differences in the dust scale heights from the azimuthal intensity variation in the right panel of Figure 16.

To summarize the discussion above, a small deprojected beam is desirable, i.e., a small observed beam and aligned position angles of the beam and ring. Especially if the beam is large and the position angles of the beam and ring are misaligned, the beam smearing can cause azimuthal intensity variations even if the ring is geometrically thin. Since the beam smearing reduces the intensity along the ridge, we should convolve model images with the beam before comparing them with the observation.

Please wait… references are loading.
10.3847/1538-4357/abe5a6