Rossby Wave Instability and Substructure Formation in 3D Non-Ideal MHD Wind-Launching Disks

Chun-Yen Hsu,1 Zhi-Yun Li,1 Yisheng Tu,1 Xiao Hu,1,2 Min-Kai Lin,3,4
1Department of Astronomy, University of Virginia, Charlottesville, VA 22904, USA
2Department of Astronomy, University of Florida, Gainesville, FL 32608, USA
3Institute of Astronomy and Astrophysics, Academia Sinica, Taipei 10617, Taiwan
4Physics Division, National Center for Theoretical Sciences, Taipei 10617, Taiwan
E-mail: kdj8qp@virginia.edu
(Accepted XXX. Received YYY; in original form ZZZ)
Abstract

Rings and gaps are routinely observed in the dust continuum emission of protoplanetary discs (PPDs). How they form and evolve remains debated. Previous studies have demonstrated the possibility of spontaneous gas rings and gaps formation in wind-launching disks. Here, we show that such gas substructures are unstable to the Rossby Wave Instability (RWI) through numerical simulations. Specifically, shorter wavelength azimuthal modes develop earlier, and longer wavelength ones dominate later, forming elongated (arc-like) anti-cyclonic vortices in the rings and (strongly magnetized) cyclonic vortices in the gaps that persist until the end of the simulation. Highly elongated vortices with aspect ratios of 10 or more are found to decay with time in our non-ideal MHD simulation, in contrast with the hydro case. This difference could be caused by magnetically induced motions, particularly strong meridional circulations with large values of the azimuthal component of the vorticity, which may be incompatible with the columnar structure preferred by vortices. The cyclonic and anti-cyclonic RWI vortices saturate at moderate levels, modifying but not destroying the rings and gaps in the radial gas distribution of the disk. In particular, they do not shut off the poloidal magnetic flux accumulation in low-density regions and the characteristic meridional flow patterns that are crucial to the ring and gap formation in wind-launching disks. Nevertheless, the RWI and their associated vortices open up the possibility of producing non-axisymmetric dust features observed in a small fraction of protoplanetary disks through non-ideal MHD, although detailed dust treatment is needed to explore this possibility.

keywords:
accretion, accretion discs – MHD – protoplanetary discs – instabilities
pubyear: 2015pagerange: Rossby Wave Instability and Substructure Formation in 3D Non-Ideal MHD Wind-Launching DisksB

1 Introduction

The detection of substructures is a watershed event in protoplanetary disk research (PPDs, e.g., ALMA Partnership et al., 2015; Andrews et al., 2018). While axisymmetric rings, gaps, and cavities are the most common features, non-axisymmetric structures such as spiral arms, lobes, and arcs are also observed. A particular type of non-axisymmetric feature is vortices, which are detected in, e.g., HD 142527 using CO emission lines (Boehler et al., 2021) and possibly through near-infrared scattered light (Marr & Dong, 2022). Several mechanisms are proposed to produce the substructures, including planet-disc interactions (e.g., Dong et al., 2015; Bae et al., 2017), sintering of volatile ices outside snow lines (Okuzumi et al., 2016), sharp change of disk properties at the dead zone boundary (e.g., Flock et al., 2015; Ruge et al., 2016), secular gravitational instability (e.g., Takahashi & Inutsuka, 2016; Tominaga et al., 2020), zonal flows arising in magnetorotational instability (MRI) turbulence (e.g., Johansen et al., 2009; Krapp et al., 2018), and non-ideal MHD effects and magnetic disk winds (e.g., Suriano et al., 2017; Riols & Lesur, 2019; Cui & Bai, 2021).

Magnetic fields have been observed in dense star-forming cores of molecular clouds that collapse to form stars and circumstellar disks (e.g., Maury et al., 2018; Galametz et al., 2018; Le Gouellec et al., 2020). The magnetic field is widely believed to drive the disk evolution, particularly through a magnetized disk wind (e.g., Bai & Stone, 2013; Lesur et al., 2022), which removes angular momentum through a laminar torque. Two-dimensional (2D) non-ideal magnetohydrodynamic (MHD) numerical simulations have shown that axisymmetric disks may evolve into stable rings and gaps in wind-launching disks (e.g., Suriano et al., 2017, 2018). More recent 2D non-ideal MHD simulations included detailed chemistry networks in computing the non-ideal MHD coefficients, finding that stable rings and gaps still form robustly, as long as the dimensionless Elsasser number ΛΛ\Lambdaroman_Λ (which measures the degree of field-neutral coupling) is of order unity or larger (e.g., Nolan et al., 2023). The question that naturally arises is: are such axisymmetric rings and gaps stable in three dimensions (3D), particularly to the Rossby wave instability (RWI, Lovelace et al., 1999; Li et al., 2000)?

The RWI occurs when the local Rossby wave is trapped in the non-self-gravitating disks. It may allow non-axisymmetric perturbations to grow into well-defined vortices (e.g., Zaqarashvili et al., 2021). The RWI has been investigated in the 2D (vertically integrated) linear theories (e.g., Lovelace et al., 1999; Li et al., 2000; Umurhan, 2010; Ono et al., 2016) and numerical simulations (e.g., Li et al., 2001; Varnière & Tagger, 2006; Lyra et al., 2008; Ono et al., 2018; Cimerman & Rafikov, 2023). In the 3D disk (where the vertical structure is accounted for), the RWI has also been studied using both linear theories (e.g., Lin, 2012, 2013) and numerical simulations (e.g., Lyra & Mac Low, 2012; Richard et al., 2013), mostly with a prescribed initially axisymmetric substructure that is observationally motivated but not formed self-consistently. The spontaneous formation of rings and gaps in 2D (axisymmetric) non-ideal MHD simulations (e.g., Suriano et al., 2017, 2018; Nolan et al., 2023) provides a rare opportunity to examine the RWI of axisymmetric structures produced through consistent dynamics using 3D simulations. Just as importantly, such 3D simulations also allow us to investigate the formation of disk substructures and the development of RWI in them simultaneously. In this paper, we seek to determine whether the non-linear development of RWI in the rings and gaps formed in the non-ideal MHD wind-launching disks can lead to the formation of vortices and, if so, how they affect the rings and gaps in the radial gas distribution. Vortices are important to investigate because of their potential for dust trapping (see, e.g., a recent review by Bae et al., 2023), which may facilitate planetesimal formation.

This paper is organized as follows. We present our simulation setup in §2, including the governing equations. §3 describes our chemistry model for computing the abundances of charged species, which are used to determine the non-ideal MHD coefficients. In §4, we show results of both 2D and 3D simulations, focusing on RWI and its effects on the disk substructure formation. Our results are briefly discussed and summarized in §5.

2 Simulation setup

We use Athena++ (Stone et al., 2020) to solve the non-ideal MHD equations:

ρt+(ρ𝐕)=0,𝜌𝑡𝜌𝐕0\frac{\partial\rho}{\partial t}+\nabla\cdot(\rho{\rm\bf V})=0,divide start_ARG ∂ italic_ρ end_ARG start_ARG ∂ italic_t end_ARG + ∇ ⋅ ( italic_ρ bold_V ) = 0 , (1)
(ρ𝐕)t+(ρ𝐕𝐕+P𝐁𝐁4π)=ρΦ,𝜌𝐕𝑡𝜌𝐕𝐕superscriptP𝐁𝐁4𝜋𝜌Φ\frac{\partial(\rho{\rm\bf V})}{\partial t}+\nabla\cdot(\rho{\rm\bf V}{\rm\bf V% }+\rm P^{*}-\frac{{\rm\bf B}{\rm\bf B}}{4\pi})=-\rho\nabla\Phi,divide start_ARG ∂ ( italic_ρ bold_V ) end_ARG start_ARG ∂ italic_t end_ARG + ∇ ⋅ ( italic_ρ bold_VV + roman_P start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT - divide start_ARG bold_BB end_ARG start_ARG 4 italic_π end_ARG ) = - italic_ρ ∇ roman_Φ , (2)
edt+[(ed+P)𝐕𝐁(𝐁𝐕)4π+1c(ηO𝐉+ηAD𝐉)]subscript𝑒𝑑𝑡delimited-[]subscript𝑒𝑑superscriptP𝐕𝐁𝐁𝐕4𝜋1csubscript𝜂O𝐉subscript𝜂ADsubscript𝐉perpendicular-to\displaystyle\frac{\partial e_{d}}{\partial t}+\nabla\cdot\left[(e_{d}+\rm P^{% *}){\rm\bf V}-\frac{{\rm\bf B}({\rm\bf B}\cdot{\rm\bf V})}{4\pi}+\frac{1}{c}% \left(\eta_{O}{\rm\bf J}+\eta_{\rm AD}{\rm\bf J}_{\perp}\right)\right]divide start_ARG ∂ italic_e start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_t end_ARG + ∇ ⋅ [ ( italic_e start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT + roman_P start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) bold_V - divide start_ARG bold_B ( bold_B ⋅ bold_V ) end_ARG start_ARG 4 italic_π end_ARG + divide start_ARG 1 end_ARG start_ARG roman_c end_ARG ( italic_η start_POSTSUBSCRIPT roman_O end_POSTSUBSCRIPT bold_J + italic_η start_POSTSUBSCRIPT roman_AD end_POSTSUBSCRIPT bold_J start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ) ]
=ρ(𝐕Φ)Λc,absent𝜌𝐕ΦsubscriptΛc\displaystyle=-\rho({\rm\bf V}\cdot\nabla\Phi)-\Lambda_{\rm c},= - italic_ρ ( bold_V ⋅ ∇ roman_Φ ) - roman_Λ start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT , (3)

and the induction equation

𝐁t+×(𝐕×𝐁)4πc×[ηO𝐉+ηAD𝐉],𝐁𝑡𝐕𝐁4𝜋𝑐delimited-[]subscript𝜂𝑂𝐉subscript𝜂ADsubscript𝐉perpendicular-to\frac{\partial{\rm\bf B}}{\partial t}+\nabla\times({\rm\bf V}\times{\rm\bf B})% -\frac{4\pi}{c}\nabla\times\left[\eta_{O}{\rm\bf J}+\eta_{\rm AD}{\rm\bf J}_{% \perp}\right],divide start_ARG ∂ bold_B end_ARG start_ARG ∂ italic_t end_ARG + ∇ × ( bold_V × bold_B ) - divide start_ARG 4 italic_π end_ARG start_ARG italic_c end_ARG ∇ × [ italic_η start_POSTSUBSCRIPT italic_O end_POSTSUBSCRIPT bold_J + italic_η start_POSTSUBSCRIPT roman_AD end_POSTSUBSCRIPT bold_J start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ] , (4)

where ρ𝜌\rhoitalic_ρ and 𝐕𝐕{\rm\bf V}bold_V are gas mass density and velocity, 𝐁𝐁{\rm\bf B}bold_B is the magnetic field, P=P+B2/(8π)superscriptPPsuperscriptB28𝜋\rm P^{*}=P+B^{2}/(8\pi)roman_P start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = roman_P + roman_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ( 8 italic_π ) is the total (thermal [P𝑃Pitalic_P] and magnetic) pressure, Φ=GM/rΦ𝐺𝑀𝑟\Phi=-GM/rroman_Φ = - italic_G italic_M / italic_r is the gravitational potential of the central star, ed=ρV2/2+P/(γ1)+B2/(8π)subscript𝑒𝑑𝜌superscript𝑉22𝑃𝛾1superscript𝐵28𝜋e_{d}=\rho V^{2}/2+{P}/(\gamma-1)+B^{2}/(8\pi)italic_e start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT = italic_ρ italic_V start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 + italic_P / ( italic_γ - 1 ) + italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ( 8 italic_π ) is the energy density, γ𝛾\gammaitalic_γ is the adiabatic index, 𝐉𝐉{\rm\bf J}bold_J is the current density, 𝐉=𝐁×(𝐉×𝐁)/(B2)subscript𝐉perpendicular-to𝐁𝐉𝐁superscript𝐵2{\rm\bf J}_{\perp}={\rm\bf B}\times({\rm\bf J}\times{\rm\bf B})/(B^{2})bold_J start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT = bold_B × ( bold_J × bold_B ) / ( italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) is the component of 𝐉𝐉{\rm\bf J}bold_J perpendicular to the magnetic field, ηOsubscript𝜂𝑂\eta_{O}italic_η start_POSTSUBSCRIPT italic_O end_POSTSUBSCRIPT and ηADsubscript𝜂AD\eta_{\rm AD}italic_η start_POSTSUBSCRIPT roman_AD end_POSTSUBSCRIPT are the Ohmic and ambipolar diffusivities, and ΛcsubscriptΛc\Lambda_{\rm c}roman_Λ start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT is the cooling term.

2.1 Simulation domain

We use spherical coordinates (r,θ,ϕ𝑟𝜃italic-ϕr,\theta,\phiitalic_r , italic_θ , italic_ϕ) to perform the 2D (axisymmetric, with quantities independent of ϕitalic-ϕ\phiitalic_ϕ) and 3D simulations. The simulation domain goes from 1 to 316 AU in the radial direction, 0.05 to π𝜋\piitalic_π-0.05 in the polar direction, and, for 3D simulations, from 0 to 2π2𝜋2\pi2 italic_π in the azimuthal direction. We use a logarithmic grid in the radial direction, with 80 base cells and a ratio of 1.07416 for the sizes of two adjacent cells. The grid is uniform in the polar direction, with 96 cells at the root level. For 3D simulations, we use 32 base cells in the azimuthal direction. Three levels of static mesh refinement (SMR) are used, with the grid size changing by a factor of 2 between adjacent levels, as indicated in Fig. 1. Specifically, the finest level extends from 10 to 100 au in the radial direction and about 2.5 scale heights (similar-to\sim 0.13 radians) below to above the mid-plane. The disc scale height is resolved by about 12.6 grids. The second and third finest levels cover, respectively, 5.57 to 176 au and 3.14 to 316 au radially and a polar region within 0.25 radians and 0.5 radians of the mid-plane.

Refer to caption
Figure 1: The inner part of the 3D grid, showing three levels of static mesh refinement.

2.2 Boundary conditions

We use modified outflow boundary conditions at both inner and outer radial boundaries. At the inner boundary, instead of copying gas density and pressure from the innermost active zone to ghost zones, we apply the same power law used in the gas initialization to extend gas density and pressure to ghost zones. For gas velocity, the azimuthal velocity follows a Keplerian curve, and the other two components are copied from the innermost active zone while preventing mass from outside from entering the simulation domain. Reflective conditions are used for the θ𝜃\thetaitalic_θ boundaries, and periodic boundary conditions are applied in the azimuthal direction.

2.3 Initial conditions

We perform two types of simulations: (i) 2D simulations with an initial hydrostatic equilibrium state and a magnetic field generated from vector potential to ensure a divergence-free B field (i.e., 𝐁=0𝐁0\nabla\cdot{\rm\bf B}=0∇ ⋅ bold_B = 0) and (ii) 3D simulations that restart from 2D simulations either at relatively late times or near the beginning.

We adopt the initial temperature and density profiles used in Hu et al. (2022). Specifically, we divide the simulation domain into a cold, dense disk and a hot, low-density corona, with a constant aspect ratio, h/r=0.05𝑟0.05h/r=0.05italic_h / italic_r = 0.05, where hhitalic_h is the disk scale height. We limit the cold, dense portion to two scale heights above (and below) the midplane, i.e., in the region π/2θ0<θ<π/2+θ0𝜋2subscript𝜃0𝜃𝜋2subscript𝜃0\pi/2-\theta_{0}<\theta<\pi/2+\theta_{0}italic_π / 2 - italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT < italic_θ < italic_π / 2 + italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, where θ0=arctan(2h/r)subscript𝜃02𝑟\theta_{0}=\arctan{(2h/r)}italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = roman_arctan ( 2 italic_h / italic_r ). The gas density and temperature in the disk midplane both follow a power law with index p=1.5𝑝1.5p=-1.5italic_p = - 1.5 and q=1𝑞1q=-1italic_q = - 1, respectively:

ρ(r,π/2)=ρ0(r/r0)p𝜌𝑟𝜋2subscript𝜌0superscript𝑟subscript𝑟0𝑝\displaystyle\rho(r,\pi/2)=\rho_{0}(r/r_{0})^{p}italic_ρ ( italic_r , italic_π / 2 ) = italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_r / italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT italic_p end_POSTSUPERSCRIPT (5)
T(r,π/2)=T0(r/r0)q𝑇𝑟𝜋2subscript𝑇0superscript𝑟subscript𝑟0𝑞\displaystyle T(r,\pi/2)=T_{0}(r/r_{0})^{q}italic_T ( italic_r , italic_π / 2 ) = italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_r / italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT italic_q end_POSTSUPERSCRIPT

where r0=1subscript𝑟01r_{0}=1italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 1 au is the radius of the inner boundary of the computational domain, and ρ0=2.667×1010g/cm3subscript𝜌02.667superscript1010gsuperscriptcm3\rho_{0}=2.667\times 10^{-10}{\rm g/cm^{-3}}italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 2.667 × 10 start_POSTSUPERSCRIPT - 10 end_POSTSUPERSCRIPT roman_g / roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT and T0=570subscript𝑇0570T_{0}=570italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 570 K are the density and temperature at r0subscript𝑟0r_{0}italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. To ensure a smooth transition between the cold disk and hot corona, we adopt the following vertical profile for the temperature:

T(r,θ)={T(r,π/2)if |θπ/2|<θ0T(r,π/2)exp[(|θπ/2|θ0)/θ0×ln(160)]if θ0|θπ/2|2θ0160T(r,π/2);if |θπ/2|>2θ0T(r,\theta)=\begin{cases}T(r,\pi/2)&\text{if }|\theta-\pi/2|<\theta_{0}\\ T(r,\pi/2)\ exp[(|\theta-\pi/2|\\ \ -\theta_{0})/\theta_{0}\times\ln(160)]&\text{if }\theta_{0}\leq|\theta-\pi/2% |\leq 2\theta_{0}\\ 160\ T(r,\pi/2);&\text{if }|\theta-\pi/2|>2\theta_{0}\\ \end{cases}italic_T ( italic_r , italic_θ ) = { start_ROW start_CELL italic_T ( italic_r , italic_π / 2 ) end_CELL start_CELL if | italic_θ - italic_π / 2 | < italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL italic_T ( italic_r , italic_π / 2 ) italic_e italic_x italic_p [ ( | italic_θ - italic_π / 2 | end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL - italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) / italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT × roman_ln ( 160 ) ] end_CELL start_CELL if italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≤ | italic_θ - italic_π / 2 | ≤ 2 italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL 160 italic_T ( italic_r , italic_π / 2 ) ; end_CELL start_CELL if | italic_θ - italic_π / 2 | > 2 italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_CELL end_ROW (6)

We use a quick βcoolsubscript𝛽cool\beta_{\rm cool}italic_β start_POSTSUBSCRIPT roman_cool end_POSTSUBSCRIPT cooling scheme with a cooling timescale of only 1010superscript101010^{-10}10 start_POSTSUPERSCRIPT - 10 end_POSTSUPERSCRIPT of the local orbital period, so the temperature profile is effectively fixed over time. The vertical density profile is generated based on hydrostatic equilibrium, i.e.,vr=vθ=0subscript𝑣𝑟subscript𝑣𝜃0v_{r}=v_{\theta}=0italic_v start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT = italic_v start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT = 0.

The initial magnetic field is computed from the magnetic vector potential used in Zanni et al. (2007):

Br(r,θ)=1r2sinθAϕθ,subscript𝐵𝑟𝑟𝜃1superscript𝑟2sin𝜃subscript𝐴italic-ϕ𝜃\displaystyle B_{r}(r,\theta)=\frac{1}{r^{2}{\rm sin}\theta}\frac{\partial A_{% \phi}}{\partial\theta},italic_B start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ( italic_r , italic_θ ) = divide start_ARG 1 end_ARG start_ARG italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_sin italic_θ end_ARG divide start_ARG ∂ italic_A start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_θ end_ARG , (7)
Bθ(r,θ)=1rsinθAϕr,subscript𝐵𝜃𝑟𝜃1𝑟sin𝜃subscript𝐴italic-ϕ𝑟\displaystyle B_{\theta}(r,\theta)=-\frac{1}{r\ {\rm sin}\theta}\frac{\partial A% _{\phi}}{\partial r},italic_B start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT ( italic_r , italic_θ ) = - divide start_ARG 1 end_ARG start_ARG italic_r roman_sin italic_θ end_ARG divide start_ARG ∂ italic_A start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_r end_ARG , (8)
Bϕ=0,subscript𝐵italic-ϕ0\displaystyle B_{\phi}=0,italic_B start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT = 0 , (9)

with

Aϕ(r,θ)=43r02Bp,0(rsinθr0)341(1+2cot2θ)5/8subscript𝐴italic-ϕ𝑟𝜃43superscriptsubscript𝑟02subscript𝐵p0superscript𝑟sin𝜃subscript𝑟0341superscript12cosuperscriptt2𝜃58\displaystyle A_{\phi}(r,\theta)=\frac{4}{3}r_{0}^{2}B_{\rm p,0}\left(\frac{r{% \rm sin}\theta}{r_{0}}\right)^{\frac{3}{4}}\frac{1}{(1+2{\rm cot^{2}\theta})^{% 5/8}}italic_A start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT ( italic_r , italic_θ ) = divide start_ARG 4 end_ARG start_ARG 3 end_ARG italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_B start_POSTSUBSCRIPT roman_p , 0 end_POSTSUBSCRIPT ( divide start_ARG italic_r roman_sin italic_θ end_ARG start_ARG italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT divide start_ARG 3 end_ARG start_ARG 4 end_ARG end_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG ( 1 + 2 roman_c roman_o roman_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_θ ) start_POSTSUPERSCRIPT 5 / 8 end_POSTSUPERSCRIPT end_ARG (10)

where Bp,0subscript𝐵p0B_{\rm p,0}italic_B start_POSTSUBSCRIPT roman_p , 0 end_POSTSUBSCRIPT sets the scale for the poloidal field strength. The magnetic field setup is the same as that of Bai & Stone (2017), Suriano et al. (2018), and Hu et al. (2022). In all our simulations, Bp,0subscript𝐵p0B_{\rm p,0}italic_B start_POSTSUBSCRIPT roman_p , 0 end_POSTSUBSCRIPT is set by plasma β=103𝛽superscript103\beta=10^{3}italic_β = 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT in the mid-plane.

We perform two types of 3D simulations to investigate (1) whether axisymmetric substructures formed in 2D simulations are stable to RWI and (2) how disc substructures and RWI develop simultaneously. For the first type, we restart in 3D a 2D simulation after it has evolved for a certain amount of time (typically tst=2000subscript𝑡st2000t_{\rm st}=2000italic_t start_POSTSUBSCRIPT roman_st end_POSTSUBSCRIPT = 2000 years) when prominent rings and gaps have formed. For the second type, we restart from the beginning of the simulation (tst=0subscript𝑡st0t_{\rm st}=0italic_t start_POSTSUBSCRIPT roman_st end_POSTSUBSCRIPT = 0). Since the focus of the 3D simulations is on RWI, we add a random perturbation to the radial velocity up to 10%percent\%% of the local sound speed at the time of restart to seed the instability.

3 Non-ideal MHD Effects

3.1 Chemical network for charge abundances

To determine the non-ideal MHD coefficients in equation (4), the abundances of charged species must be computed first. We use a reduced chemical network similar to those of Umebayashi & Nakano (1990) and Nishi et al. (1991). Specifically, we include the following elements: H, He, C, O, and Mg. Table. 2 in the Appendix lists their initial abundances.

We include cosmic ray ionization, which dominates the ionization in the bulk of the disk. We adopt a total ionization rate (including ionization of H2 and He, Umebayashi & Nakano, 1990) of ξ=1017s1𝜉superscript1017superscripts1\xi=10^{-17}{\rm s^{-1}}italic_ξ = 10 start_POSTSUPERSCRIPT - 17 end_POSTSUPERSCRIPT roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT. The rate equation for species i𝑖iitalic_i of the fractional abundance, xini/nHsubscript𝑥𝑖subscript𝑛𝑖subscript𝑛Hx_{i}\equiv n_{i}/n_{\rm H}italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ≡ italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT / italic_n start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT, can be written as:

1nHdxidt=1nHjkcrxj+j,kkjkxjxkximki,mxm,1subscript𝑛Hdsubscript𝑥𝑖d𝑡1subscript𝑛Hsubscript𝑗subscript𝑘crsubscript𝑥𝑗subscript𝑗𝑘subscript𝑘𝑗𝑘subscript𝑥𝑗subscript𝑥𝑘subscript𝑥𝑖subscript𝑚subscript𝑘𝑖𝑚subscript𝑥𝑚\displaystyle\frac{1}{n_{\rm H}}\frac{{\rm d}x_{i}}{{\rm d}t}=\frac{1}{n_{\rm H% }}\sum_{j}k_{\rm cr}x_{j}+\sum_{j,k}k_{jk}x_{j}x_{k}-x_{i}\sum_{m}k_{i,m}x_{m},divide start_ARG 1 end_ARG start_ARG italic_n start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT end_ARG divide start_ARG roman_d italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG roman_d italic_t end_ARG = divide start_ARG 1 end_ARG start_ARG italic_n start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT roman_cr end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT + ∑ start_POSTSUBSCRIPT italic_j , italic_k end_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT italic_j italic_k end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT - italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT italic_i , italic_m end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT , (11)

where kcrsubscript𝑘crk_{\rm cr}italic_k start_POSTSUBSCRIPT roman_cr end_POSTSUBSCRIPT is the cosmic-ray ionization rate, and kjksubscript𝑘𝑗𝑘k_{jk}italic_k start_POSTSUBSCRIPT italic_j italic_k end_POSTSUBSCRIPT and ki,msubscript𝑘𝑖𝑚k_{i,m}italic_k start_POSTSUBSCRIPT italic_i , italic_m end_POSTSUBSCRIPT are the formation and destruction rates by two-body reactions, respectively.

We include three types of reactions in the paper: gas-phase, gas-grain, and grain-grain. Table 3 lists all gas-phase reaction rates used in the paper. The gas-phase reaction rates are adapted from UMIST database McElroy et al. (2013), with the rate coefficient represented by:

k(T)=αc(T300K)βcexp(γcT),𝑘𝑇subscript𝛼csuperscript𝑇300Ksubscript𝛽cexpsubscript𝛾cT\displaystyle k(T)=\alpha_{\rm c}\left(\frac{T}{300\rm{K}}\right)^{\beta_{\rm c% }}\rm{exp}\left(-\frac{\gamma_{\rm c}}{T}\right),italic_k ( italic_T ) = italic_α start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT ( divide start_ARG italic_T end_ARG start_ARG 300 roman_K end_ARG ) start_POSTSUPERSCRIPT italic_β start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT end_POSTSUPERSCRIPT roman_exp ( - divide start_ARG italic_γ start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT end_ARG start_ARG roman_T end_ARG ) , (12)

where αcsubscript𝛼c\alpha_{\rm c}italic_α start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT is the constant factor, βcsubscript𝛽c\beta_{\rm c}italic_β start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT is the power-index of the temperature dependence, and γcsubscript𝛾c\gamma_{\rm c}italic_γ start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT measures the energy barrier of the reaction.

We adopt a power law of n(a)a3.5proportional-to𝑛𝑎superscript𝑎3.5n(a)\varpropto a^{-3.5}italic_n ( italic_a ) ∝ italic_a start_POSTSUPERSCRIPT - 3.5 end_POSTSUPERSCRIPT for the grain size distribution, with the power index of the standard MRN distribution (Mathis et al., 1977). There are 30 size bins logarithmically distributed from the minimum aminsubscript𝑎mina_{\rm min}italic_a start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT to the maximum amaxsubscript𝑎maxa_{\rm max}italic_a start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT. A grain material density of 3 g/cm3superscriptcm3{\rm cm}^{3}roman_cm start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT is adopted. The dust-to-gas mass ratio is set to 0.01. The gas-grain and grain-grain reaction rates are calculated as average reaction rates following Grassi et al. (2019):

kc,g(T)=aminamaxφ(a)kc,g(a,T)daaminamaxφ(a)da,delimited-⟨⟩subscript𝑘𝑐𝑔𝑇superscriptsubscriptsubscript𝑎minsubscript𝑎max𝜑𝑎subscript𝑘𝑐𝑔𝑎𝑇differential-d𝑎superscriptsubscriptsubscript𝑎minsubscript𝑎max𝜑𝑎differential-d𝑎\displaystyle\langle k_{c,g}(T)\rangle=\frac{\int_{a_{\rm min}}^{a_{\rm max}}% \varphi(a)k_{c,g}(a,T){\rm d}a}{\int_{a_{\rm min}}^{a_{\rm max}}\varphi(a){\rm d% }a},⟨ italic_k start_POSTSUBSCRIPT italic_c , italic_g end_POSTSUBSCRIPT ( italic_T ) ⟩ = divide start_ARG ∫ start_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_φ ( italic_a ) italic_k start_POSTSUBSCRIPT italic_c , italic_g end_POSTSUBSCRIPT ( italic_a , italic_T ) roman_d italic_a end_ARG start_ARG ∫ start_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_φ ( italic_a ) roman_d italic_a end_ARG , (13)
kg,g(T)=aminamaxaminamaxφ(a)kg,g(a,a,T)φ(a)dadaaminamaxaminamaxφ(a)φ(a)dadadelimited-⟨⟩subscript𝑘𝑔𝑔𝑇superscriptsubscriptsubscript𝑎minsubscript𝑎maxsuperscriptsubscriptsubscript𝑎minsubscript𝑎max𝜑𝑎subscript𝑘𝑔𝑔𝑎superscript𝑎𝑇𝜑superscript𝑎differential-d𝑎differential-dsuperscript𝑎superscriptsubscriptsubscript𝑎minsubscript𝑎maxsuperscriptsubscriptsubscript𝑎minsubscript𝑎max𝜑𝑎𝜑superscript𝑎differential-d𝑎differential-dsuperscript𝑎\displaystyle\langle k_{g,g}(T)\rangle=\frac{\int_{a_{\rm min}}^{a_{\rm max}}% \int_{a_{\rm min}}^{a_{\rm max}}\varphi(a)k_{g,g}(a,a^{{}^{\prime}},T)\varphi(% a^{{}^{\prime}}){\rm d}a{\rm d}a^{{}^{\prime}}}{\int_{a_{\rm min}}^{a_{\rm max% }}\int_{a_{\rm min}}^{a_{\rm max}}\varphi(a)\varphi(a^{{}^{\prime}}){\rm d}a{% \rm d}a^{{}^{\prime}}}⟨ italic_k start_POSTSUBSCRIPT italic_g , italic_g end_POSTSUBSCRIPT ( italic_T ) ⟩ = divide start_ARG ∫ start_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ∫ start_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_φ ( italic_a ) italic_k start_POSTSUBSCRIPT italic_g , italic_g end_POSTSUBSCRIPT ( italic_a , italic_a start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT , italic_T ) italic_φ ( italic_a start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT ) roman_d italic_a roman_d italic_a start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT end_ARG start_ARG ∫ start_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ∫ start_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_φ ( italic_a ) italic_φ ( italic_a start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT ) roman_d italic_a roman_d italic_a start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT end_ARG (14)

where kc,g(a,T)subscript𝑘𝑐𝑔𝑎𝑇k_{c,g}(a,T)italic_k start_POSTSUBSCRIPT italic_c , italic_g end_POSTSUBSCRIPT ( italic_a , italic_T ) denotes the rate coefficient for gas-grain reactions, and kg,g(a,a,T)subscript𝑘𝑔𝑔𝑎superscript𝑎𝑇k_{g,g}(a,a^{{}^{\prime}},T)italic_k start_POSTSUBSCRIPT italic_g , italic_g end_POSTSUBSCRIPT ( italic_a , italic_a start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT , italic_T ) is the grain-grain interaction rate coefficient. To simplify the symbol, we note as=ai+ajsubscript𝑎ssubscript𝑎𝑖subscript𝑎𝑗a_{\rm s}=a_{i}+a_{j}italic_a start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT = italic_a start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT as the sum of the grain sizes for the grain-grain interaction. We assume a negligible radius for the gas-phase species compared to the grains, so as=ajsubscript𝑎ssubscript𝑎𝑗a_{\rm s}=a_{j}italic_a start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT = italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT when we calculate kc,g(a,T)subscript𝑘𝑐𝑔𝑎𝑇k_{c,g}(a,T)italic_k start_POSTSUBSCRIPT italic_c , italic_g end_POSTSUBSCRIPT ( italic_a , italic_T ). Following Nishi et al. (1991) and Grassi et al. (2019), we consider up to two elemental charges sticking on the grains (e.g., Z=0,±1,±2𝑍0plus-or-minus1plus-or-minus2Z=0,\pm 1,\pm 2italic_Z = 0 , ± 1 , ± 2). Following Draine & Sutin (1987); Zhao et al. (2018), we consider three cases for kc,g(a,T)subscript𝑘𝑐𝑔𝑎𝑇k_{c,g}(a,T)italic_k start_POSTSUBSCRIPT italic_c , italic_g end_POSTSUBSCRIPT ( italic_a , italic_T ) and kg,g(a,a,T)subscript𝑘𝑔𝑔𝑎superscript𝑎𝑇k_{g,g}(a,a^{{}^{\prime}},T)italic_k start_POSTSUBSCRIPT italic_g , italic_g end_POSTSUBSCRIPT ( italic_a , italic_a start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT , italic_T ):

ki,j0(as,T)=πas2(8kBTπmr)1/2[1+(πe22askBT)1/2]S(T,Zj),subscriptsuperscript𝑘0𝑖𝑗subscript𝑎𝑠𝑇𝜋superscriptsubscript𝑎𝑠2superscript8subscript𝑘B𝑇𝜋subscript𝑚r12delimited-[]1superscript𝜋superscript𝑒22subscript𝑎𝑠subscript𝑘B𝑇12𝑆𝑇subscript𝑍𝑗\displaystyle k^{0}_{i,j}(a_{s},T)=\pi a_{s}^{2}\left(\frac{8k_{\rm B}T}{\pi m% _{\rm r}}\right)^{1/2}\left[1+\left(\frac{\pi e^{2}}{2a_{s}k_{\rm B}T}\right)^% {1/2}\right]S(T,Z_{j}),italic_k start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT ( italic_a start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT , italic_T ) = italic_π italic_a start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( divide start_ARG 8 italic_k start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT italic_T end_ARG start_ARG italic_π italic_m start_POSTSUBSCRIPT roman_r end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT [ 1 + ( divide start_ARG italic_π italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_a start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT italic_T end_ARG ) start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT ] italic_S ( italic_T , italic_Z start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) , (15)
ki,j(as,T)=πas2(8kBTπmr)1/2(1ZjZie2askBT)subscriptsuperscript𝑘𝑖𝑗subscript𝑎𝑠𝑇𝜋superscriptsubscript𝑎𝑠2superscript8subscript𝑘B𝑇𝜋subscript𝑚r121subscript𝑍𝑗subscript𝑍𝑖superscript𝑒2subscript𝑎𝑠subscript𝑘B𝑇\displaystyle k^{-}_{i,j}(a_{s},T)=\pi a_{s}^{2}\left(\frac{8k_{\rm B}T}{\pi m% _{\rm r}}\right)^{1/2}\left(1-\frac{Z_{j}}{Z_{i}}\frac{e^{2}}{a_{s}k_{\rm B}T}\right)italic_k start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT ( italic_a start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT , italic_T ) = italic_π italic_a start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( divide start_ARG 8 italic_k start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT italic_T end_ARG start_ARG italic_π italic_m start_POSTSUBSCRIPT roman_r end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT ( 1 - divide start_ARG italic_Z start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG start_ARG italic_Z start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG divide start_ARG italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_a start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT italic_T end_ARG )
[1+(2askBTe22ZjZi)1/2]S(T,Zj),absentdelimited-[]1superscript2subscript𝑎𝑠subscript𝑘B𝑇superscript𝑒22subscript𝑍𝑗subscript𝑍𝑖12𝑆𝑇subscript𝑍𝑗\displaystyle\cdot\left[1+\left(\frac{2}{\frac{a_{s}k_{\rm B}T}{e^{2}}-2\frac{% Z_{j}}{Z_{i}}}\right)^{1/2}\right]S(T,Z_{j}),⋅ [ 1 + ( divide start_ARG 2 end_ARG start_ARG divide start_ARG italic_a start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT italic_T end_ARG start_ARG italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG - 2 divide start_ARG italic_Z start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG start_ARG italic_Z start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG end_ARG ) start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT ] italic_S ( italic_T , italic_Z start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) , (16)
ki,j+(as,T)=πas2(8kBTπmr)1/2subscriptsuperscript𝑘𝑖𝑗subscript𝑎𝑠𝑇𝜋superscriptsubscript𝑎𝑠2superscript8subscript𝑘B𝑇𝜋subscript𝑚r12\displaystyle k^{+}_{i,j}(a_{s},T)=\pi a_{s}^{2}\left(\frac{8k_{\rm B}T}{\pi m% _{\rm r}}\right)^{1/2}italic_k start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT ( italic_a start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT , italic_T ) = italic_π italic_a start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( divide start_ARG 8 italic_k start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT italic_T end_ARG start_ARG italic_π italic_m start_POSTSUBSCRIPT roman_r end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT
[1+(4askBTe2+3ZjZi)1/2]2exp(θνe2askBT)S(T,Zj),absentsuperscriptdelimited-[]1superscript4subscript𝑎𝑠subscript𝑘B𝑇superscript𝑒23subscript𝑍𝑗subscript𝑍𝑖122expsubscript𝜃𝜈superscript𝑒2subscript𝑎𝑠subscript𝑘B𝑇𝑆𝑇subscript𝑍𝑗\displaystyle\cdot\left[1+\left(\frac{4a_{s}k_{\rm B}T}{e^{2}}+3\frac{Z_{j}}{Z% _{i}}\right)^{-1/2}\right]^{2}{\rm exp}\left(-\frac{\theta_{\nu}e^{2}}{a_{s}k_% {\rm B}T}\right)S(T,Z_{j}),⋅ [ 1 + ( divide start_ARG 4 italic_a start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT italic_T end_ARG start_ARG italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + 3 divide start_ARG italic_Z start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG start_ARG italic_Z start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT - 1 / 2 end_POSTSUPERSCRIPT ] start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_exp ( - divide start_ARG italic_θ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_a start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT italic_T end_ARG ) italic_S ( italic_T , italic_Z start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) , (17)

where e𝑒eitalic_e is the elemental charge, mr=mimj/(mi+mj)subscript𝑚rsubscript𝑚𝑖subscript𝑚𝑗subscript𝑚𝑖subscript𝑚𝑗m_{\rm r}=m_{i}m_{j}/(m_{i}+m_{j})italic_m start_POSTSUBSCRIPT roman_r end_POSTSUBSCRIPT = italic_m start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_m start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT / ( italic_m start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + italic_m start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) is the reduced mass, S(T)𝑆𝑇S(T)italic_S ( italic_T ) is the sticking coefficient, and θν=Zj3/2/[Zi(Zi+Zj)]subscript𝜃𝜈superscriptsubscript𝑍𝑗32delimited-[]subscript𝑍𝑖subscript𝑍𝑖subscript𝑍𝑗\theta_{\nu}=Z_{j}^{3/2}/[Z_{i}(\sqrt{Z_{i}}+\sqrt{Z_{j}})]italic_θ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT = italic_Z start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 / 2 end_POSTSUPERSCRIPT / [ italic_Z start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( square-root start_ARG italic_Z start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG + square-root start_ARG italic_Z start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG ) ]. ki,j0subscriptsuperscript𝑘0𝑖𝑗k^{0}_{i,j}italic_k start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT is for charged species colliding with neutral grains (e.g., ZiZj=0subscript𝑍𝑖subscript𝑍𝑗0Z_{i}Z_{j}=0italic_Z start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_Z start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = 0). ki,jsubscriptsuperscript𝑘𝑖𝑗k^{-}_{i,j}italic_k start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT is the rate for opposite charge reactions (e.g., ZiZj<0subscript𝑍𝑖subscript𝑍𝑗0Z_{i}Z_{j}<0italic_Z start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_Z start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT < 0). ki,j+subscriptsuperscript𝑘𝑖𝑗k^{+}_{i,j}italic_k start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT is the reaction rate between species with the same electrical properties (e.g., ZiZj>0subscript𝑍𝑖subscript𝑍𝑗0Z_{i}Z_{j}>0italic_Z start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_Z start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT > 0). If the collisions are between charged species and grains, then the index i𝑖iitalic_i and j𝑗jitalic_j represent the charged species and grains, respectively. The index i𝑖iitalic_i and j𝑗jitalic_j represent the lighter and heavier grains for grain-grain interactions. The sticking coefficients S(T,Z)𝑆𝑇𝑍S(T,Z)italic_S ( italic_T , italic_Z ) of electrons on grains are taken from Grassi et al. (2019, their Table 3). Following Umebayashi & Nakano (1990), we set the sticking coefficients for non-electron charged species on grains to S(T,Z)=1𝑆𝑇𝑍1S(T,Z)=1italic_S ( italic_T , italic_Z ) = 1.

3.2 Non-ideal MHD Coefficients

Once the charge abundances are known, the ambipolar, Hall, and Ohmic diffusivities can be computed using the standard procedure based on the parallel, Pedersen, and Hall conductivities (e.g., Wardle, 2007; Pinto et al., 2008):

σO=ecBΣjnj|Zj|βj,n,subscript𝜎O𝑒𝑐𝐵subscriptΣ𝑗subscript𝑛𝑗subscript𝑍𝑗subscript𝛽𝑗𝑛\displaystyle\sigma_{\rm O}=\frac{ec}{B}\Sigma_{j}n_{j}|Z_{j}|\beta_{j,n},italic_σ start_POSTSUBSCRIPT roman_O end_POSTSUBSCRIPT = divide start_ARG italic_e italic_c end_ARG start_ARG italic_B end_ARG roman_Σ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT | italic_Z start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT | italic_β start_POSTSUBSCRIPT italic_j , italic_n end_POSTSUBSCRIPT , (18)
σH=ecBΣjnj|Zj|βj,n21+βj,n2,subscript𝜎H𝑒𝑐𝐵subscriptΣ𝑗subscript𝑛𝑗subscript𝑍𝑗superscriptsubscript𝛽𝑗𝑛21superscriptsubscript𝛽𝑗𝑛2\displaystyle\sigma_{\rm H}=-\frac{ec}{B}\Sigma_{j}\frac{n_{j}|Z_{j}|\beta_{j,% n}^{2}}{1+\beta_{j,n}^{2}},italic_σ start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT = - divide start_ARG italic_e italic_c end_ARG start_ARG italic_B end_ARG roman_Σ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT divide start_ARG italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT | italic_Z start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT | italic_β start_POSTSUBSCRIPT italic_j , italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 1 + italic_β start_POSTSUBSCRIPT italic_j , italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG , (19)
σP=ecBΣjnj|Zj|βj,n1+βj,n2subscript𝜎P𝑒𝑐𝐵subscriptΣ𝑗subscript𝑛𝑗subscript𝑍𝑗subscript𝛽𝑗𝑛1superscriptsubscript𝛽𝑗𝑛2\displaystyle\sigma_{\rm P}=\frac{ec}{B}\Sigma_{j}\frac{n_{j}|Z_{j}|\beta_{j,n% }}{1+\beta_{j,n}^{2}}italic_σ start_POSTSUBSCRIPT roman_P end_POSTSUBSCRIPT = divide start_ARG italic_e italic_c end_ARG start_ARG italic_B end_ARG roman_Σ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT divide start_ARG italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT | italic_Z start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT | italic_β start_POSTSUBSCRIPT italic_j , italic_n end_POSTSUBSCRIPT end_ARG start_ARG 1 + italic_β start_POSTSUBSCRIPT italic_j , italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG (20)

where

βj,n=eZjBmjcmj+mnρnRj,n(T)subscript𝛽𝑗𝑛𝑒subscript𝑍𝑗𝐵subscript𝑚𝑗𝑐subscript𝑚𝑗subscript𝑚𝑛subscript𝜌𝑛subscript𝑅𝑗𝑛𝑇\displaystyle\beta_{j,n}=\frac{eZ_{j}B}{m_{j}c}\frac{m_{j}+m_{n}}{\rho_{n}R_{j% ,n}(T)}italic_β start_POSTSUBSCRIPT italic_j , italic_n end_POSTSUBSCRIPT = divide start_ARG italic_e italic_Z start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_B end_ARG start_ARG italic_m start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_c end_ARG divide start_ARG italic_m start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT + italic_m start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT end_ARG start_ARG italic_ρ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT italic_j , italic_n end_POSTSUBSCRIPT ( italic_T ) end_ARG (21)

is the Hall parameter for collisions between the j𝑗jitalic_j-th charged species and neutral hydrogen, njsubscript𝑛𝑗n_{j}italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT is the number density of the j𝑗jitalic_j-th species, ρnsubscript𝜌𝑛\rho_{n}italic_ρ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT and mnsubscript𝑚𝑛m_{n}italic_m start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT are the neutral (molecular) hydrogen mass density and mass, respectively, and Rj,n(T)subscript𝑅𝑗𝑛𝑇R_{j,n}(T)italic_R start_POSTSUBSCRIPT italic_j , italic_n end_POSTSUBSCRIPT ( italic_T ) is the momentum exchange rate coefficient, which is discussed in detail in Appendix B.

From the conductivities, we can compute the Ohmic, Hall, and ambipolar diffusivities as follows:

ηO=c24πσO,subscript𝜂Osuperscript𝑐24𝜋subscript𝜎O\displaystyle\eta_{\rm O}=\frac{c^{2}}{4\pi\sigma_{\rm O}},italic_η start_POSTSUBSCRIPT roman_O end_POSTSUBSCRIPT = divide start_ARG italic_c start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 4 italic_π italic_σ start_POSTSUBSCRIPT roman_O end_POSTSUBSCRIPT end_ARG , (22)
ηH=c24πσσHσ,subscript𝜂Hsuperscript𝑐24𝜋subscript𝜎perpendicular-tosubscript𝜎Hsubscript𝜎perpendicular-to\displaystyle\eta_{\rm H}=\frac{c^{2}}{4\pi\sigma_{\perp}}\frac{\sigma_{\rm H}% }{\sigma_{\perp}},italic_η start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT = divide start_ARG italic_c start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 4 italic_π italic_σ start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT end_ARG divide start_ARG italic_σ start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT end_ARG start_ARG italic_σ start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT end_ARG , (23)
ηAD=c24πσσPσηO,subscript𝜂ADsuperscript𝑐24𝜋subscript𝜎perpendicular-tosubscript𝜎Psubscript𝜎perpendicular-tosubscript𝜂O\displaystyle\eta_{\rm AD}=\frac{c^{2}}{4\pi\sigma_{\perp}}\frac{\sigma_{\rm P% }}{\sigma_{\perp}}-\eta_{\rm O},italic_η start_POSTSUBSCRIPT roman_AD end_POSTSUBSCRIPT = divide start_ARG italic_c start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 4 italic_π italic_σ start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT end_ARG divide start_ARG italic_σ start_POSTSUBSCRIPT roman_P end_POSTSUBSCRIPT end_ARG start_ARG italic_σ start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT end_ARG - italic_η start_POSTSUBSCRIPT roman_O end_POSTSUBSCRIPT , (24)

where σ=σH2+σP2subscript𝜎perpendicular-tosuperscriptsubscript𝜎H2superscriptsubscript𝜎P2\sigma_{\perp}=\sqrt{\sigma_{\rm H}^{2}+\sigma_{\rm P}^{2}}italic_σ start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT = square-root start_ARG italic_σ start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_σ start_POSTSUBSCRIPT roman_P end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG.

The dimensionless Elsasser numbers for these diffusivities are given by ΛO=VA0ΩηO,ΛH=VA0ΩηHformulae-sequencesubscriptΛOsubscript𝑉A0Ωsubscript𝜂OsubscriptΛHsubscript𝑉A0Ωsubscript𝜂H\Lambda_{\rm O}=\frac{V_{\rm A0}}{\Omega\eta_{\rm O}},\ \Lambda_{\rm H}=\frac{% V_{\rm A0}}{\Omega\eta_{\rm H}}roman_Λ start_POSTSUBSCRIPT roman_O end_POSTSUBSCRIPT = divide start_ARG italic_V start_POSTSUBSCRIPT A0 end_POSTSUBSCRIPT end_ARG start_ARG roman_Ω italic_η start_POSTSUBSCRIPT roman_O end_POSTSUBSCRIPT end_ARG , roman_Λ start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT = divide start_ARG italic_V start_POSTSUBSCRIPT A0 end_POSTSUBSCRIPT end_ARG start_ARG roman_Ω italic_η start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT end_ARG and ΛAD=VA0ΩηADsubscriptΛADsubscript𝑉A0Ωsubscript𝜂AD\Lambda_{\rm AD}=\frac{V_{\rm A0}}{\Omega\eta_{\rm AD}}roman_Λ start_POSTSUBSCRIPT roman_AD end_POSTSUBSCRIPT = divide start_ARG italic_V start_POSTSUBSCRIPT A0 end_POSTSUBSCRIPT end_ARG start_ARG roman_Ω italic_η start_POSTSUBSCRIPT roman_AD end_POSTSUBSCRIPT end_ARG, where VA0=B02/(4πρ0)subscript𝑉A0superscriptsubscript𝐵024𝜋subscript𝜌0V_{\rm A0}=B_{0}^{2}/(4\pi\rho_{0})italic_V start_POSTSUBSCRIPT A0 end_POSTSUBSCRIPT = italic_B start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ( 4 italic_π italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) is the Alfvén speed, with the lower subscript ’0’ denoting the local quantity at each location, and ΩΩ\Omegaroman_Ω is the Keplerian angular velocity.

To mimic the better magnetic coupling (and thus reduced diffusivity) due to higher ionization levels expected in the lower density regions near the disk surface and in the disk wind, we follow Suriano et al. (2018) and multiply the magnetic diffusivities in Eq. (22) to (24) by the following θ𝜃\thetaitalic_θ dependence:

f(θ)={exp(cos2(θ+θ0)2(h/r)2)ifθ<π2θ01ifπ2θ0θπ2+θ0exp(cos2(θθ0)2(h/r)2)ifθ>π2+θ0.𝑓𝜃casesexpsuperscriptcos2𝜃subscript𝜃02superscript𝑟2if𝜃𝜋2subscript𝜃01if𝜋2subscript𝜃0𝜃𝜋2subscript𝜃0expsuperscriptcos2𝜃subscript𝜃02superscript𝑟2if𝜃𝜋2subscript𝜃0\displaystyle f(\theta)=\begin{cases}{\rm exp}\left(-\frac{{\rm cos}^{2}(% \theta+\theta_{0})}{2(h/r)^{2}}\right)&\mbox{if}\quad\theta<\frac{\pi}{2}-% \theta_{0}\\ 1&\mbox{if}\quad\frac{\pi}{2}-\theta_{0}\leq\theta\leq\frac{\pi}{2}+\theta_{0}% \\ {\rm exp}\left(-\frac{{\rm cos}^{2}(\theta-\theta_{0})}{2(h/r)^{2}}\right)&% \mbox{if}\quad\theta>\frac{\pi}{2}+\theta_{0}\end{cases}.italic_f ( italic_θ ) = { start_ROW start_CELL roman_exp ( - divide start_ARG roman_cos start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_θ + italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) end_ARG start_ARG 2 ( italic_h / italic_r ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) end_CELL start_CELL if italic_θ < divide start_ARG italic_π end_ARG start_ARG 2 end_ARG - italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL 1 end_CELL start_CELL if divide start_ARG italic_π end_ARG start_ARG 2 end_ARG - italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≤ italic_θ ≤ divide start_ARG italic_π end_ARG start_ARG 2 end_ARG + italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL roman_exp ( - divide start_ARG roman_cos start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_θ - italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) end_ARG start_ARG 2 ( italic_h / italic_r ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) end_CELL start_CELL if italic_θ > divide start_ARG italic_π end_ARG start_ARG 2 end_ARG + italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_CELL end_ROW . (25)

4 Formation of Gas Rings and Gaps in 2D and Their Stability in 3D

Table 1: Models
Name Description Starting time tstsubscript𝑡stt_{\rm st}italic_t start_POSTSUBSCRIPT roman_st end_POSTSUBSCRIPT
S-2D Shu-type power-law ion abundance 0
T-2D chemical network 0
T-3D-2000 chemical network 2000 years
T-3D-0 chemical network 0

We will mainly consider models with the charge abundances computed from the chemical network discussed in § 3.1. The abundances are shown in Appendix A (Fig. 15) for an MRN-type grain size distribution from amin=0.5μsubscript𝑎min0.5𝜇a_{\rm min}=0.5\ \muitalic_a start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT = 0.5 italic_μm to amax=25μsubscript𝑎max25𝜇a_{\rm max}=25\ \muitalic_a start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 25 italic_μm. These abundances are saved as a look-up table and referenced during the simulation to compute the non-ideal MHD coefficients in each computational cell. These models are labeled by the letter “T " (for “Table") in Table 1, which lists all 2D and 3D models discussed in the paper. For comparison, we also consider a 2D model with the simple power-law prescription for the dominant molecular ion from Shu (1992) that was used in previous work along a similar line (Suriano et al., 2018; Hu et al., 2022). This model is denoted by the letter “S" (for “Shu") in Table 1. The initial Elsasser numbers for these two model types are shown in Fig. 16 of Appendix A for reference.

4.1 2D Simulations

Refer to caption
Figure 2: Rings and gaps in 2D Models S-2D (top panels) and T-2D (bottom panels) at a representative time t=2000𝑡2000t=2000italic_t = 2000 years in the inner disk (up to 35 au in radius). Plotted in panels (a) and (c) are the mass density distribution (color map), velocity streamlines (black lines with arrows), and magnetic field lines (red lines with arrows) on a meridional plane. A magnetized disk wind is clearly visible in both cases. Panels (b) and (d) display the surface density of the disk normalized to its initial value, highlighting the formation of prominent rings and gaps, especially in Model T-2D.

We start with a comparison of the 2D (axisymmetric) simulations using the Shu-type power-law prescription (Model S-2D) and the more realistic charge abundances from a look-up table computed from a chemical network including dust grains (Model T-2D). In Fig. 2, we plot a meridional (left panels) and face-on (right panels) view of the simulations at a representative time t=2000𝑡2000t=2000italic_t = 2000 years when stable rings and gaps are formed. We use cylindrical coordinates (R, ϕitalic-ϕ\phiitalic_ϕ, z) in the meridional plots. The rings and gaps in Model T-2D are more prominent than those in Model S-2D because the gas in the former is better coupled to the magnetic field than in the latter in the bulk of the disk material (see Fig. 16). Our results add weight to the conclusion based on previous work in the literature (e.g., Bai & Stone, 2013; Lesur, 2021), especially that of Nolan et al. (2023), that ring and gap formation is a robust phenomenon in 2D that is not sensitive to the detailed treatment of the ambipolar diffusivity as long as the disk gas is reasonably well coupled to the magnetic field, with an ambipolar Elsasser number ΛΛ\Lambdaroman_Λ of order unity or larger. The question arises: Are the rings (and gaps) formed in 2D stable in 3D, particularly regarding Rayleigh instability and Rossby Wave Instability (RWI)?

Refer to caption
Figure 3: Radial profiles of (a) the surface density of the disk normalized to its initial value, (b) the function (r)𝑟\mathcal{L}(r)caligraphic_L ( italic_r ) (defined in equation [26]) that controls the RWI for adiabatic flows, (c) the function (r)𝑟\mathcal{H}(r)caligraphic_H ( italic_r ) (defined in equation [27] that controls the RWI for locally isothermal flows, and (d) the specific angular momentum on the midplane, for the disks with rings and gaps shown in Fig. 2. Note the local maximum and the local minimum in the (r)𝑟\mathcal{L}(r)caligraphic_L ( italic_r ) and (r)𝑟\mathcal{H}(r)caligraphic_H ( italic_r ) profiles in panels (b) and (c) are the positions where the disk may be unstable to the RWI. Regions with specific angular momentum increasing with radius are stable to the Rayleigh instability for rotating flows.

One way to check the Rossby wave stability is to compute the quantity (r)𝑟\mathcal{L}(r)caligraphic_L ( italic_r ) defined in Lovelace et al. (1999) as a function of radius:

(r)(r)S2/γ(r).𝑟𝑟superscriptS2𝛾𝑟\mathcal{L}(r)\equiv\mathcal{F}(r){\rm S}^{2/\gamma}(r).caligraphic_L ( italic_r ) ≡ caligraphic_F ( italic_r ) roman_S start_POSTSUPERSCRIPT 2 / italic_γ end_POSTSUPERSCRIPT ( italic_r ) . (26)

The result is shown in Fig. 3. Panel (a) shows the radial profiles of the column density of Model S-2D and Model T-2D at the time shown in Fig. 2. The column density contrast between the rings and gaps in Model T-2D is higher than in Model S-2D, consistent with Fig. 2. Panel (b) of Fig. 3 shows the radial profiles of the quantity (r)𝑟\mathcal{L}(r)caligraphic_L ( italic_r ) of Model S-2D and Model T-2D, where (r)Σ/(2ωz)𝑟Σ2subscript𝜔𝑧\mathcal{F}(r)\approx\Sigma/(2\omega_{z})caligraphic_F ( italic_r ) ≈ roman_Σ / ( 2 italic_ω start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ), ωzz^(×𝐕)subscript𝜔𝑧^𝑧𝐕\omega_{z}\equiv\hat{z}\cdot(\nabla\times{\rm\bf V})italic_ω start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ≡ over^ start_ARG italic_z end_ARG ⋅ ( ∇ × bold_V ), and SP/ΣγS𝑃superscriptΣ𝛾{\rm S}\equiv P/\Sigma^{\gamma}roman_S ≡ italic_P / roman_Σ start_POSTSUPERSCRIPT italic_γ end_POSTSUPERSCRIPT with P𝑃Pitalic_P denoting the vertically integrated pressure and γ𝛾\gammaitalic_γ is the adiabatic index. It is formally derived for an adiabatic flow for razor-thin disks, which differs from our case, where strong cooling keeps the flow nearly isothermal locally. In this case, the relevant quantity to evaluate is the inverse of a generalized vortensity, defined as (Lin, 2012):

(r)2ΩΣcs2κ2=Σcs2r(rvϕ)/r=Σcs2ωz,𝑟2ΩΣsuperscriptsubscript𝑐𝑠2superscript𝜅2Σsuperscriptsubscript𝑐𝑠2𝑟𝑟subscript𝑣italic-ϕ𝑟Σsuperscriptsubscript𝑐𝑠2subscript𝜔𝑧\mathcal{H}(r)\equiv\frac{2\Omega\Sigma c_{s}^{2}}{\kappa^{2}}=\frac{\Sigma c_% {s}^{2}r}{\partial(rv_{\phi})/\partial r}=\frac{\Sigma c_{s}^{2}}{\omega_{z}},caligraphic_H ( italic_r ) ≡ divide start_ARG 2 roman_Ω roman_Σ italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_κ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG = divide start_ARG roman_Σ italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_r end_ARG start_ARG ∂ ( italic_r italic_v start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT ) / ∂ italic_r end_ARG = divide start_ARG roman_Σ italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_ω start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT end_ARG , (27)

where κ𝜅\kappaitalic_κ is the local epicycle frequency. This quantity is plotted in panel (c). In addition, we plot in panel (d) the distribution of the specific angular momentum rvϕ𝑟subscript𝑣italic-ϕr\ v_{\phi}italic_r italic_v start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT on the disk midplane as a function of radius r𝑟ritalic_r. A few (limited) regions have specific angular momentum decreasing with radius; they are prone to Rayleigh instability, which may quickly develop to limit the amplitude of the pressure gradient-induced deviation from the background Keplerian rotation in 3D (e.g., Ono et al., 2018).

Lovelace et al. (1999) and Lin (2012) found that a local extremum of (r)𝑟\mathcal{L}(r)caligraphic_L ( italic_r ) and (r)𝑟\mathcal{H}(r)caligraphic_H ( italic_r ) is a necessary criterion for destabilizing the Rossby waves for adiabatic and locally isothermal flows, respectively. Panels (b) and (c) of Figure 3 show local extrema are ubiquitous in our case. It indicates that the Rossby waves might grow throughout our simulation domain if there is an initial azimuthal variation to trigger them. However, our simulations do not strictly follow the situation envisioned in Lovelace et al. (1999) or Lin (2012) because of the inclusion of additional physical effects such as the magnetic field, non-ideal MHD effects, and disk winds. In what follows, we will determine whether the rings are indeed unstable to RWI through direct 3D simulations.

4.2 Stability of 2D Rings and Gaps in 3D

Refer to caption
Figure 4: Evolution of 2D rings and gaps in 3D. Plotted in the first column are the surface density distribution normalized to its initial value at three representative times t=2340𝑡2340t=2340italic_t = 2340 (panel a), 2980298029802980 (panel e), and 3490349034903490 yr (panel i) of Model T-3D-2000 restarted from Model T-2D at t=2000𝑡2000t=2000italic_t = 2000 yr. Similarly, the second, third, and fourth columns show, respectively, the midplane mass density normalized to its initial value, the midplane plasma-βzsubscript𝛽𝑧\beta_{z}italic_β start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT based on the vertical field component Bzsubscript𝐵𝑧B_{z}italic_B start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT alone, and the midplane radial velocity normalized by the sound speed. An animated version of the figure can be found on the website: %{\color{red}␣[Please␣add␣a␣vertical␣line␣with␣easily␣visible␣horizontal␣tick␣marks␣to␣all␣panels␣in␣the␣second␣column␣since␣the␣discussion␣in␣the␣text␣often␣refers␣to␣rings␣at␣one␣radius␣(e.g.,␣11␣au)␣or␣another␣(e.g.,␣20␣au);␣there␣is␣no␣need␣for␣the␣white␣space␣surrounding␣the␣labeled␣numbers␣(it␣is␣there␣because␣of␣the␣way␣I␣made␣the␣figure␣using␣PPT)␣and␣keep␣only␣10,␣20,␣and␣30␣au␣labels␣(i.e.,␣remove␣5,␣15,␣25,␣and␣35␣labels␣to␣avoid␣crowding␣but␣make␣sure␣that␣the␣tick␣marks␣at␣5,␣15,␣25,␣and␣35␣are␣clearly␣visible␣as␣at␣10,␣20,␣and␣30).␣Make␣sure␣it␣is␣easy␣for␣the␣reader␣to␣follow␣the␣discussion␣in␣the␣text␣with␣such␣added␣lines␣and␣that␣the␣discussion␣is␣consistent␣with␣the␣added␣lines.␣Remove␣the␣vertical␣line␣in␣panel␣l␣since␣it␣is␣no␣longer␣needed.␣Please␣do␣the␣same␣for␣Fig.␣10]}https://virginia.box.com/s/b1wxk512g1xzm371thk9h7w4l1sfl8qk

To check the stability of the disk substructures formed in the 2D simulations, we restart Model T-2D (which produced more prominent substructures than the other 2D model, Model S-2D) from the representative time shown in Fig. 2. We include a random perturbation to the radial component of the velocity up to 10%percent1010\%10 % of the local sound speed to seed any potential instability that may grow. The results of the restarted 3D model, T-3D-2000, are shown in Fig. 4. The figure plots the surface density distribution normalized to its initial value at the beginning of the 2D simulation (t=0𝑡0t=0italic_t = 0), the midplane mass density normalized to its initial value, the midplane plasma-βzsubscript𝛽𝑧\beta_{z}italic_β start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT based on the vertical magnetic field component, and the midplane radial velocity normalized by the local sound speed at three representative times: t=2340𝑡2340t=2340italic_t = 2340 (top panels), 2980298029802980 (middle), and 3490349034903490 years (bottom panels). The link to an animated version of the figure can be found in its caption.

Figure 4 and its associated animation clearly show the development of non-axisymmetric structures, which follow the patterns expected for RWI; we will focus primarily on the rings since their azimuthal variations show up more clearly in the surface and mass density maps (RWI-induced structures in the gaps will be discussed towards the end of the next subsection). Specifically, the instability first develops in the inner disk and progressively moves to larger radii with longer local dynamical times. For a given ring of roughly constant radius, the azimuthal mode number m𝑚mitalic_m of the dominant non-axisymmetric features decreases with time. For example, at the relatively early time of t=2340𝑡2340t=2340italic_t = 2340 yr (or 340 yr after the 3D restart), there are 8 well-defined over-dense clumps in the midplane mass density distribution that are regularly spaced on the ring at 11similar-toabsent11\sim 11∼ 11 au (panel b), even though the clumps are less visible in the surface density at the same time (panel a). The m=8𝑚8m=8italic_m = 8 mode is even more prominent in the distribution of the midplane plasma-βzsubscript𝛽𝑧\beta_{z}italic_β start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT, which is much higher inside the over-dense clumps than between them (panel c). The same mode also shows up in the midplane radial velocity distribution, with the fastest radial outward motion occurring between the over-dense clumps and at a slightly smaller radius (panel d). We note that the time t=2340𝑡2340t=2340italic_t = 2340 yr corresponds to about 9 times the local orbital period at the unstable ring location, comparable to the saturation time for the fiducial model studied by Ono et al. (2018).

Similar patterns are observed at the later time t=2980𝑡2980t=2980italic_t = 2980 yr but for a ring at a larger radius of 20similar-toabsent20\sim 20∼ 20 au (compared to 11similar-toabsent11\sim 11∼ 11 au discussed in the last paragraph). Here, the m=8𝑚8m=8italic_m = 8 mode is the most prominent in the radial velocity distribution (panel h), with 8 regularly spaced regions of fast inflow at approximately half of the local sound speed. The fast-inflow regions are located slightly outside the visibly perturbed 20 au-radius ring, with the inflow apparently displacing the ring inward, producing a regularly spaced undulation in the radial direction, which is also evident in the surface density distribution (panel e). The undulation soon leads to the formation of 8 over-dense regions similar to those for the inner ring near 11 au at the earlier time t=2340𝑡2340t=2340italic_t = 2340 yr (see panel b), as can be seen from the animated version of the figure (see times around t=3000𝑡3000t=3000italic_t = 3000 yr). There is, therefore, little doubt that the 20 au-ring is RWI unstable, with a well-developed m=8𝑚8m=8italic_m = 8 mode at t=2980𝑡2980t=2980italic_t = 2980 yr, corresponding to about 11 local orbital periods, similar to the 11 au ring at a comparable local dynamical time (normalized by the orbital period). The mode is less evident in the midplane plasma-βzsubscript𝛽𝑧\beta_{z}italic_β start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT distribution because it has more radial substructures than the density distribution, which makes the high plasma-βzsubscript𝛽𝑧\beta_{z}italic_β start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ring near 20 au stand out less than the high mass density ring at a similar radius.

The lack of one-to-one correspondence between the substructures in the plasma-βzsubscript𝛽𝑧\beta_{z}italic_β start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT and density distributions is illustrated by the faint ring inside the prominent gap around 10 au, which is barely visible in the density map (panel f) but stands out clearly as a high plasma-βzsubscript𝛽𝑧\beta_{z}italic_β start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ring against the lower plasma-βzsubscript𝛽𝑧\beta_{z}italic_β start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT adjacent gaps (panel g). This low-density but high βzsubscript𝛽𝑧\beta_{z}italic_β start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ring is the 11 au ring showing prominent m=8𝑚8m=8italic_m = 8 azimuthal mode at the earlier time t=2340𝑡2340t=2340italic_t = 2340 yr (panel b). At this earlier time, at least seven well-defined rings were located inside the 20 au ring discussed at the later time t=2980𝑡2980t=2980italic_t = 2980 yr, with four outside the prominent 10 au gap and at least three inside. Visible azimuthal structures develop earlier in these (inner) rings than in the 20 au ring, with higher order modes at earlier times. By t=2980𝑡2980t=2980italic_t = 2980 yr, only relatively low-order azimuthal modes (with m𝑚mitalic_m of a few) are clearly visible in the density map (panel f). In particular, two prominent arcs (corresponding to the m=2𝑚2m=2italic_m = 2 azimuthal mode) develop in the ring immediately interior to the 10 au ring around t=2700𝑡2700t=2700italic_t = 2700 yr (see the animation of the density map), which has been smeared into a long arc with moderate azimuthal density variation by the time shown in panel (f). The trend for lower-order modes to dominate later is consistent with the expected evolution and saturation of RWI modes (see, e.g., Ono et al., 2018).

The trend is vividly illustrated by the last frame of the simulation when the surface density and midplane density of the 20 au ring become dominated by two well-defined, relatively short, over-dense arcs (panels i and j) at t=3490𝑡3490t=3490italic_t = 3490 yr, corresponding to approximately 17 local orbital periods after the 3D restart. By this time, the azimuthal variations of the midplane density of the rings at larger radii remain dominated by higher m𝑚mitalic_m modes. In particular, an m=7𝑚7m=7italic_m = 7 mode starts to show up clearly in the outermost ring (at 32similar-toabsent32\sim 32∼ 32 au) displayed at the time of panel (j), corresponding to approximately 8 local orbital periods, similar to the times (normalized by the local orbit period) when the m=8𝑚8m=8italic_m = 8 mode becomes prominent for the inner 11 au (panel b) and 20 au (panel f) rings.

It is unclear how long the low-order m=2𝑚2m=2italic_m = 2 mode for the 20 au ring at the last frame would survive. A hint for the longer-term evolution of RWI-induced structures comes from the rings at relatively small radii, which evolved for longer local dynamical times than their larger-radius counterparts. For example, two prominent over-dense (relatively long) arcs develop in the 7.5 au ring around 2650265026502650 yr (approximately 31 local orbital periods after the 3D restart), and last until 2760276027602760 yr for 110similar-toabsent110\sim 110∼ 110 yr or 5similar-toabsent5\sim 5∼ 5 local orbital periods (see the animated version of the midplane density). A single (m=1𝑚1m=1italic_m = 1) over-dense arc dominates the appearance of the ring at some of the later times (e.g., t=3300𝑡3300t=3300italic_t = 3300 yr), in agreement with the expectations based on the simpler hydro case (e.g., Ono et al., 2018). However, this is not the case at other times, e.g., at the last frame shown in panel (j), where at least two, possibly three, arcs with moderate density enhancements exist. This deviation from the expectations is unsurprising since our simulated disk is more complex than the simplest hydro case (with, e.g., non-ideal MHD and disk wind). In particular, the rings and gaps in our simulation are highly dynamic structures that can gain or lose mass and angular momentum through spatially dependent mass accretion and (magnetized) outflow.

Refer to caption
Figure 5: The time evolution of the mode amplitudes of the azimuthal surface density variation for three representative rings at radii of R18similar-to𝑅18R\sim 18italic_R ∼ 18 (panel a), 8 (b), and 5 au (c) of Model T-3D-2000. The time is normalized by their local period. Panel (a) shows the m=8𝑚8m=8italic_m = 8 mode first dominates the 18 au ring before being overtaken by the m=2𝑚2m=2italic_m = 2 mode. Panels (b) and (c) illustrate a similar trend: the higher-order modes develop first before being overtaken by low-order modes, as expected for RWI modes.

To quantify the late-time evolution of different azimuthal modes, we plot in Fig. 5 the mode amplitudes of the azimuthal surface density variations of three representative rings at approximately 5, 8, and 18 au as a function of time normalized by their local orbital period. The low-order m=2𝑚2m=2italic_m = 2 mode (see the red curve) starts to dominate at a time between 15similar-toabsent15\sim 15∼ 15 and 30similar-toabsent30\sim 30∼ 30 local orbital periods; its further evolution is truncated by the termination of the simulation for the 18 au ring (panel a). For the 8 au ring, the m=2𝑚2m=2italic_m = 2 mode decays after 10similar-toabsent10\sim 10∼ 10 local orbital periods and is replaced by the even lower-order m=1𝑚1m=1italic_m = 1 mode with a higher peak amplitude (panel b). The m=1𝑚1m=1italic_m = 1 mode decays after 30similar-toabsent30\sim 30∼ 30 orbits. Its longer-term evolution is unclear because of the simulation time limit, but a hint is offered by the 5 au ring, which reached 130similar-toabsent130\sim 130∼ 130 orbits. In this case, the m=1𝑚1m=1italic_m = 1 mode is more important than the m=2𝑚2m=2italic_m = 2 mode most of the time after 65similar-toabsent65\sim 65∼ 65 orbits, but the m=2𝑚2m=2italic_m = 2 remains significant and occasionally surpasses the m=1𝑚1m=1italic_m = 1 mode in amplitude. A caveat is that the 5 au ring is on a coarser grid than the other two rings and is resolved azimuthally by only 64 cells (see the statically refined grid in Fig. 1). The relatively low resolution may have affected the growth and saturation of the modes. Nevertheless, it is clear that, as expected for RWI, low-order modes dominate the azimuthal variation of the surface density at late times, with relatively small amplitudes. An implication is that the non-linear development of RWI modifies but does not destroy the rings formed in our non-ideal MHD simulations.

Refer to caption
Figure 6: Time evolution of (a) the surface density distribution of Model T-2D and (b) the azimuthally averaged surface density distribution of Model T-3D-2000 as a function of radius after the 3D restart at = 2000 yr. Note the prominent rings and gaps survive in 3D despite RWI.

The effects of the RWI modes on rings are further illustrated in panel (j) of Fig. 4, which shows that the 8 au ring remains distinct from its neighboring gaps (i.e., with a significant axisymmetric m=0𝑚0m=0italic_m = 0 mode) even after a relatively large number (72similar-toabsent72\sim 72∼ 72) of local orbit periods, indicating that nonlinear development of RWI modifies rather than destroys the ring. The same is true for rings at smaller radii, such as the one around 5 au, which retains its ring-shaped structure after 133similar-toabsent133\sim 133∼ 133 local orbital periods. To demonstrate the survival of the rings and gaps more clearly, we plot in Fig. 6 the azimuthally averaged surface density distribution of Model T-3D-2000 as a function of radius after its restart at t=2000𝑡2000t=2000italic_t = 2000 yr and compare it to its 2D counterpart Model T-2D. There are some differences between the two models at late times, including a more prominent ring at 5similar-toabsent5\sim 5∼ 5 au in the 2D case and a less empty gap at 10similar-toabsent10\sim 10∼ 10 au in the 3D case. Nevertheless, the azimuthally averaged radial substructures are broadly similar in the two cases, supporting the notion that rings and gaps survive in 3D despite RWI.

4.3 RWI Vortex Structure

Refer to caption
Figure 7: Structure of the dominant vortex in the 8 au-ring at τ50similar-to𝜏50\tau\sim 50italic_τ ∼ 50. Panels (a)-(c) show the mass density normalized to its initial value, the zlimit-from𝑧z-italic_z -component vorticity normalized to its local Keplerian angular speed, and the logarithmic scale of the plasma-β𝛽\betaitalic_β on the midplane, respectively. Panels (a) and (b) include the velocity streamlines (gray lines) and the panel (c) the magnetic fields (white lines). Panels (d)-(f) are the same as (a)-(c) but on a meridional plane passing through the density peak of the 8 au ring. The brown and black contour lines are 0.6 and 1.0 contours of the mass density normalized to its initial value. An animated version of the figure can be found at https://virginia.box.com/s/499xq1qt2j9c9vgxzljlb19x610tfnqe
Refer to caption
Figure 8: Detailed structure of the dominant RWI vortex in the 8 au ring of the Model T-3D-2000 at τ50similar-to𝜏50\tau\sim 50italic_τ ∼ 50. Panels (a) and (c) show the radial profiles of the surface density normalized to its initial value and azimuthal velocity deviation from the local Keplerian value normalized to local sound speed, respectively, on a meridional plane passing through the density peak at 8 au. Panels (b) and (d) show the azimuthal profiles of the surface density normalized to its initial value and radial velocity normalized to the local sound speed of the same vortex. The red dash-dotted lines mark the locations where the mass density normalized to its initial value is 1.0 (corresponding to the black density contour in Fig. 7).

In this subsection, we analyze the structure of the vortices for the 8 au ring at the time t=3150𝑡3150t=3150italic_t = 3150 yr (or 1150 yr after the 3D restart), when the m=1𝑚1m=1italic_m = 1 azimuthal mode reaches its peak amplitude (see Fig. 5b). We will concentrate on the broad features of the vortices since their detailed structures may not be adequately resolved in our 3D global non-ideal MHD simulations. At this time, the m=2𝑚2m=2italic_m = 2 mode has a significant amplitude as well, leading to a secondary density peak next to the primary one, as shown in Fig. 7a and Fig. 8b (the one with a smaller value of the azimuthal angle ϕitalic-ϕ\phiitalic_ϕ). As expected, these vortices are anti-cyclonic, with the flow streamlines on the mid-plane circling the local density maxima in a direction opposite to the Keplerian rotation (see Fig. 7a).

The corresponding negative axial vorticity is shown in panel (b) of Fig. 7. Note that the spatial distributions of the axial vorticity in and around the vortices are patchier compared to 3D hydro simulations (e.g., Fig. 8 of Richard et al., 2013), likely because the magnetic fields included in our simulation make the disk more dynamically active and can produce spatially variable vorticity even in the absence of RWI-induced vortices. Nevertheless, regions of high negative vorticity appear relatively coherent in the vertical direction, as shown in their Rz𝑅𝑧R-zitalic_R - italic_z distribution (panel e), similar to the columnar structure seen in the hydro case (e.g., Fig. 8 of Richard et al., 2013).

The streamlines in panels (d) and (e) show that the flow on the poloidal (Rz𝑅𝑧R-zitalic_R - italic_z) plane has a downdraft in the outer part of the vortex (with larger radii than that of the density peak) and an updraft in the inner part (near the density peak), with some hint of a clockwise vortex in between. This pattern is reminiscent of the one shown in the lower-left panel of Fig. 13 of Meheut et al. (2010) (in the region between their 2.8similar-toabsent2.8\sim 2.8∼ 2.8 and 3.0ri3.0subscript𝑟𝑖3.0\ r_{i}3.0 italic_r start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT), who studied the 3D structures of hydro RWI vortices. However, the vortex is barely visible and much less prominent than the hydro case. There is a prominent counter-clockwise vortex outside the clockwise one in the hydro case, but not in our case. The difference may not be surprising since our wind-launching non-ideal MHD disk has strong meridional circulation motions, particularly in dense rings (e.g., Fig. 8 of Hu et al. 2022; see also Riols & Lesur 2019; Cui & Bai 2021 and Fig. 12d below), which can, in principle, change the dynamics and structure of the RWI vortices.

Specifically, magnetic fields may affect the motions in and around the vortices. For example, the field lines in the Rϕ𝑅italic-ϕR-\phiitalic_R - italic_ϕ plane are mostly toroidal (i.e., along ϕitalic-ϕ\phiitalic_ϕ-direction; panel c), which may resist the field bending by the cross-field motion associated with the anti-cyclonic rotation around the density peak of the vortex (panel a). However, the magnetic fields inside the over-dense vortex are relatively weak, corresponding to a typical plasma-β𝛽\betaitalic_β of order 102superscript10210^{2}10 start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (panel c), so their direct resistance to the vortex motion may be relatively modest. Nevertheless, the magnetic field and its associated outflow dominate the angular momentum transport and generate strong poloidal motions and non-Keplerian rotation throughout the disk, which could affect the RWI vortices indirectly; indeed, they give rise to the RWI-unstable substructures in the first place. However, such indirect effects are difficult to quantify.

An interesting feature of the plasma-β𝛽\betaitalic_β distribution in the Rϕ𝑅italic-ϕR-\phiitalic_R - italic_ϕ plane (panel c) is a ring of high β𝛽\betaitalic_β values offset from the density peak, where the field lines change directions abruptly. Panel (f) shows that the feature extends vertically outside the high-density part of the vortex outlined by the isodensity contours, indicating that it is primarily a feature of weak local magnetic field rather than high density. Specifically, it is primarily associated with regions where the magnetic field reversal happens, particularly in the ϕitalic-ϕ\phiitalic_ϕ-direction (i.e., where Bϕsubscript𝐵italic-ϕB_{\phi}italic_B start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT approaches zero). The field reversal in or near dense vortices is unsurprising because the gas motions there are strong enough to tangle the field lines, creating kinks where the magnetic field components (such as Bϕsubscript𝐵italic-ϕB_{\phi}italic_B start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT or Brsubscript𝐵𝑟B_{r}italic_B start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT) drop to zero. Field tangling is harder to achieve in the more strongly magnetized lower-density regions (e.g., gaps), where the field lines are straighter.

To analyze the vortices more quantitatively, we plot in Fig. 8 the radial and azimuthal profiles of the normalized surface density across the density peak of the primary vortex (panels a and b) and the radial profile of the deviation of the azimuthal velocity from the local Keplerian value (panel c) and the azimuthal profile of the radial velocity (panel d) normalized by the local sound speed. The radial density profile is asymmetric because the gap outside the vortex-forming ring is deeper than that inside (see Fig. 7 and its animated version). The super-Keplerian rotation radially interior to the density peak and sub-Keplerian rotation exterior to the peak are consistent with expectations. As expected from the counterclockwise circulation pattern of the anti-cyclonic vortex around the density peak in the Rϕ𝑅italic-ϕR-\phiitalic_R - italic_ϕ plane (see Fig. 7a), the radial velocity is positive on the larger azimuthal angle side of the density peak, reaching a maximum of 0.2similar-toabsent0.2\sim 0.2∼ 0.2 times the local sound speed. It is negative on the other side, decreasing to a minimum of only 0.1similar-toabsent0.1\sim-0.1∼ - 0.1 times the local sound speed; the asymmetry is caused by the secondary vortex nearby.

Refer to caption
Figure 9: The aspect ratio for the dominant vortex at 8 au in Model T-3D-2000 from τ=50𝜏50\tau=50italic_τ = 50 to 60606060 (the time is normalized by the local orbital period).

An important parameter to characterize a vortex is its aspect ratio χa/b𝜒𝑎𝑏\chi\equiv a/bitalic_χ ≡ italic_a / italic_b, where a𝑎aitalic_a and b𝑏bitalic_b are the major and minor axes of close streamlines (e.g., Ono et al., 2018). Because the velocity field in our non-ideal MHD simulation is much more disordered than the hydro simulations (see the velocity streamline in Fig. 7), we choose to approximate the shape of the vortex using the iso-density contour of ρ(r,ϕ)/ρ(r,t=0)=1𝜌𝑟italic-ϕ𝜌𝑟𝑡01\rho(r,\phi)/\rho(r,t=0)=1italic_ρ ( italic_r , italic_ϕ ) / italic_ρ ( italic_r , italic_t = 0 ) = 1 (i.e., the black lines for the dominant vortex in Fig. 7a). The resulting aspect ratio χ𝜒\chiitalic_χ for the dominant vortex near the 8 au radius is shown in Fig. 9 as a function of time during the period when the m=1𝑚1m=1italic_m = 1 mode dominates (50<τ<6050𝜏6050<\tau<6050 < italic_τ < 60; see panel b of Fig. 5). The aspect ratio is of order 10 or larger for the entire duration, indicating that the vortex is highly elongated, consistent with visual inspection of Fig. 7 and especially Fig. 4 and their associated animations. Lesur & Papaloizou (2009) pointed out that vortices in disks tend to be unstable to ellipsoidal instability in 3D, although Richard et al. (2013) found that highly elongated vortices with χ6similar-to𝜒6\chi\sim 6italic_χ ∼ 6 or larger can survive in 3D compressible hydro simulations. This hydro result contrasts with our non-ideal MHD simulation, where the amplitude of the dominant m=1𝑚1m=1italic_m = 1 mode at 8 au steadily decreases after peaking around τ55𝜏55\tau\approx 55italic_τ ≈ 55 despite its large aspect ratio.

There are several possibilities for the above difference. Firstly, the magnetic field in our simulation may weaken the vortex. For example, the circulating flow in the vortex must move across the predominantly toroidal magnetic field in the midplane (contrasting the flow streamlines and field lines in, e.g., Fig. 7a and c, respectively) and is expected to be resisted by magnetic tension. Lyra & Klahr (2011) found that the elliptic instability is quite destructive in magnetized disks; perhaps this explains why even elongated vortices decay in our simulations. Secondly, the vortex may be weakened by the magnetized disk wind in our simulation. Through mass and angular momentum removal, the wind induces strong meridional circulation that may be incompatible with the columnar structure preferred by the vortex. In particular, significant vertical motions are present in our simulation (see, e.g., Fig. 7d), which are different from those found in 3D hydro simulations (e.g., Meheut et al., 2010, 2012; Richard et al., 2013). Thirdly, our wind-launching non-ideal MHD disk is more dynamically active than the relatively quiescent disk envisioned in most of the previous 3D hydro RWI simulations, with an effective α𝛼\alphaitalic_α parameter of order 0.01 to 0.1 estimated from the Reynolds and Maxwell stresses, which can negatively impact the vortex’s survival (e.g., Lin, 2013).

Interestingly, many cyclonic vortices form in our restarted 3D simulation in addition to the anti-cyclonic vortices discussed above. They are formed in low-density gaps and evolve broadly like the anti-cyclonic vortices formed in dense rings: higher-order azimuthal modes dominate at earlier times and lower-order modes at later times. For example, eight cyclonic vortices (i.e., the m=8𝑚8m=8italic_m = 8 mode) are clearly visible in the similar-to\sim10-au gap from the streamlines plotted on the Rϕ𝑅italic-ϕR-\phiitalic_R - italic_ϕ plane at a relatively early time t=450𝑡450t=450italic_t = 450 years after the restart (corresponding to about 14 local orbital periods), as shown in the animated version of Fig. 7. They merge into fewer numbers at later times. In particular, at the time of t=1150𝑡1150t=1150italic_t = 1150 years shown in Fig. 7, there are only two readily identifiable cyclonic vortices with (clockwise) closed streamlines in the 10-au gap, as can be seen from panel (a) and particularly (b), which also shows that the cyclonic vortices have the largest (positive) vertical component of the vorticity ωzsubscript𝜔𝑧\omega_{z}italic_ω start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT. Besides the sign of ωzsubscript𝜔𝑧\omega_{z}italic_ω start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT, these vortices differ from the anti-cyclonic ones (with a negative ωzsubscript𝜔𝑧\omega_{z}italic_ω start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT) in one major aspect: they are more strongly magnetized, with a plasma-β𝛽\betaitalic_β typically of order unity, as can be seen from panels (c) and (f). Nevertheless, despite the strong magnetization in the gap, cyclonic vortices appear to be able to develop and survive until the end of the simulation at t=1490𝑡1490t=1490italic_t = 1490 years after the 3D restart, corresponding to similar-to\sim47 local orbital periods at 10 au.

Since the perfectly axisymmetric rings and gaps formed in 2D simulations are shown to be unstable to RWI, such idealized structures are unlikely to be produced in nature without the assumption of axisymmetry. We are thus motivated to start the 3D simulations from the very beginning to determine how the RWI interacts with the formation of the rings and gaps in the first place.

4.4 3D Simulation from the Beginning

Refer to caption
Figure 10: Simultaneous growth of rings and gaps and RWI in 3D in Model T3D-0. Plotted in the first column are the surface density distribution normalized to its initial value at three representative times t=340𝑡340t=340italic_t = 340 (panel a), 680680680680 (panel e), and 1220122012201220 yr (panel i). Similarly, the second, third, and fourth columns show, respectively, the midplane mass density normalized to its initial value, the midplane plasma-βzsubscript𝛽𝑧\beta_{z}italic_β start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT based on the vertical field component Bzsubscript𝐵𝑧B_{z}italic_B start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT alone, and the midplane radial velocity normalized by the sound speed. An animated version of the figure can be found on the website: https://virginia.box.com/s/shw9ux7dc0klb3bduj13iqvq3hj45rjq

In this section, we analyze the results of Model T-3D-0, which is the same as Model T-3D-2000 but restarts from the 2D simulation at t=0𝑡0t=0italic_t = 0 years rather than t=2000𝑡2000t=2000italic_t = 2000 years. Since the disk has no initial substructure, it is initially stable to RWI. We expect the non-axisymmetric RWI modes to grow as rings and gaps develop through non-ideal MHD processes in the wind-launching disk. The initial development of nearly axisymmetric rings and gaps is illustrated in the top panels of Fig. 10, where we plot the normalized column density, the midplane density, plasma-βzsubscript𝛽𝑧\beta_{z}italic_β start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT based on the vertical component of the magnetic field, and radial velocity normalized by the local sound speed, at the same representative time of t=340𝑡340t=340italic_t = 340 years as shown in the top panels for Model T-3D-2000 in Fig. 4. In the T-3D-2000 case, a prominent m=8𝑚8m=8italic_m = 8 azimuthal RWI mode has already become clearly visible at this time in a ring around 11similar-toabsent11\sim 11∼ 11 au radius. In the T-3D-0 case, although rings and gaps have developed near 10similar-toabsent10\sim 10∼ 10 au at this time, they remain nearly axisymmetric, likely because their amplitudes are still relatively small and it takes time for RWI modes to grow, particularly on a background ring/gap of relatively low contrast.

Non-axisymmetric RWI modes do develop at later times. For example, the middle row of Fig. 10 shows a prominent m=5𝑚5m=5italic_m = 5 mode on a ring of 9.7similar-toabsent9.7\sim 9.7∼ 9.7 au radius at the second representative time of t=680𝑡680t=680italic_t = 680 years, corresponding to about 22.5 local orbital periods. It is seen most clearly in the radial velocity map (Fig. 10h), but is also visible in the other three maps (Fig. 10efg). At later times, the non-axisymmetric structure becomes more dominated by lower-order azimuthal modes (with smaller m𝑚mitalic_m), broadly consistent with the trend expected for RWI. For example, at the end of the simulation (t=1220𝑡1220t=1220italic_t = 1220 years, shown in the bottom row), the azimuthal density perturbation is dominated by a high-density vortex (m=1𝑚1m=1italic_m = 1 mode) near the bottom of the 9.7similar-toabsent9.7\sim 9.7∼ 9.7 au-ring, although other low-order (m=2𝑚2m=2italic_m = 2 and m=3𝑚3m=3italic_m = 3) modes are also visible.

Refer to caption
Figure 11: Same as Fig. 5, but for Model T-3D-0 at R = 12 (a), 9.7 (b), and 9 au (c).

To quantify the mode evolution, we plot in Fig. 11 the time evolution of the amplitudes of various azimuthal modes, as done in Fig. 5 for the restarted model T-3D-2000. It is clear that the m=5𝑚5m=5italic_m = 5 mode initially dominates the 9.79.79.79.7 au-ring (the purple peak around τ22similar-to𝜏22\tau\sim 22italic_τ ∼ 22). As time progresses, it is first replaced by the m=3𝑚3m=3italic_m = 3 (blue), then m=2𝑚2m=2italic_m = 2 (red), and, finally, m=1𝑚1m=1italic_m = 1 (black) mode. The progression towards lower-order modes with time is also evident for the 9999 au ring shown (bottom panel), consistent with the expectation of RWI. Interestingly, the top panel shows that the azimuthal modes of the outer 12121212 au ring are dominated by m=1𝑚1m=1italic_m = 1 at relatively early times between τ13similar-to𝜏13\tau\sim 13italic_τ ∼ 13 - 16161616. This is in contrast with the outer (18181818 au) ring of Model T-3D-2000, which is dominated by the high-order m=8𝑚8m=8italic_m = 8 mode at similarly early times (see the gray curve in the top panel of Fig. 5). We attribute the different behavior to the difference in the initial disk structure when the 3D simulation starts. In the T-3D-2000 case, highly RWI unstable, well-defined rings of large contrast with their surroundings already exist at the beginning of the simulation (see the red curve in the top panel of Fig. 3). For such (unstable) conditions, the higher-order modes are expected to grow faster, as discussed earlier in Section 4.2 (see also, e.g., Ono et al. 2018). In the T-3D-0 case, there is no ring at 12 au (or anywhere else) initially. It takes time for the rings to form out of the initially smooth disk and for their amplitudes to increase relative to their surroundings (see Fig. 13 below). When the ring amplitude is smaller, the fastest-growing mode tends to be of a lower order (see, e.g., Table 1 of Ono et al., 2018), consistent with the above difference between the outer rings of Models T-3D-0 and T3D-2000.

Refer to caption
Figure 12: The structure of the dominant vortex on the 9.7 au ring in Model T-3D-0 at τ22.5similar-to𝜏22.5\tau\sim 22.5italic_τ ∼ 22.5. Panels (a) and (b) display the mass density normalized to its initial value at the midplane and the z𝑧zitalic_z component vorticity normalized to its local Keplerian period, respectively. Panels (c)-(f) plot the mass density normalized to its initial value (c), the ϕitalic-ϕ\phiitalic_ϕ component of the vorticity normalized to its local Keplerian period (d), the z𝑧zitalic_z component vorticity normalized to its local Keplerian period (e), and the logarithm of the plasma-β𝛽\betaitalic_β (f) on a meridional plane passing through the density peak of the similar-to\sim 9.2 au ring. The gray contours with arrows are velocity streamlines and the white contours (in panel [f]) with arrows are the magnetic field lines. The brown and black contour lines mark where the mass density normalized to its initial value is 0.6 and 1.0, respectively. An animated version of the figure can be found at https://virginia.box.com/s/oofe09pt2qv01wkk42vv8e0tzqg3oy7k

To illustrate the vortices formed in the T-3D-0 model more pictorially, we plot in Fig. 12 the normalized density, vertical component of the vorticity, and the plasmaβ𝛽-\beta- italic_β on the midplane and in the meridional plane passing through the center of a vortex at the same representative time of t=680𝑡680t=680italic_t = 680 yrs as in the second row of Fig. 10, similar to the maps shown Fig. 7 for model T-3D-2000. Compared to the T-3D-2000 case, the density contrast between regions of different radii is much less (compare Fig. 12a and Fig. 7), which is expected given that the T-3D-2000 model was restarted with already formed high-contrast rings and gaps. Nevertheless, regularly spaced anti-cyclonic vortices (with negative values of vertical vorticity) are clearly visible at higher-density radii (e.g., 9.3similar-toabsent9.3\sim 9.3∼ 9.3 and 10.310.310.310.3 au) and cyclonic vortices (with positive values of vertical vorticity) at lower-density radii (e.g., 8.8similar-toabsent8.8\sim 8.8∼ 8.8 and 9.89.89.89.8 au; see panels [a] and [b]), with the cyclonic vortices forming preferentially at azimuthal angles between the anti-cyclonic ones (and vice versa).

Despite the prominent vortices of both positive and negative vorticities on the Rϕ𝑅italic-ϕR-\phiitalic_R - italic_ϕ plane, the flow structure in the meridional plane at this time is rather regular, with a layer of accreting material located approximately near the midplane, which drives two prominent cells of opposite meridional circulation above and below the layer near the radius of the dominant anti-cyclonic vortex marked by the black contour in panel (c). Indeed, associated with the meridional circulations is an azimuthal component of vorticity that is larger than the vertical component, even inside the dominant RWI vortex, as seen by comparing panels (d) and (e) of Fig. 12. The strong meridional circulation, ultimately induced by the magnetic field through accretion and disk wind, is a feature absent from hydro simulations of RWI.

The meridional flow pattern in 3D is nearly identical to that studied in detail in 2D (axisymmetric) simulations (e.g., Fig.8 of Hu et al. 2022). In particular, the accretion layer is associated with regions of vanishing toroidal magnetic fields, where the field lines change directions sharply in the azimuthal direction, creating a large tension force in the (negative) azimuthal direction that brakes the rotation and drives accretion. The weak toroidal field is why the accretion layer shows up as a filament of high plasmaβ𝛽-\beta- italic_β in panel (f). The panel shows clearly that the accretion drags the poloidal field lines into a pinched configuration, which, Suriano et al. (2018) argue, leads to reconnection that lies at the heart of ring and gap formation in the first place. Clearly, the development of RWI in 3D does not shut off the mechanism for generating rings and gaps in the non-ideal MHD and wind-launching disk in the first place.

Refer to caption
Figure 13: Same as the Fig. 6, but for the Model T-3D-0 and with a different color bar.

The persistence of ring and gap formation despite the RWI is further illustrated in Fig. 13, where we compare the radial distribution of the azimuthally averaged normalized surface density at different times for Model T-3D-0 and its 2D counterpart T2D. As in the T-3D-2000 case (see Fig. 6), the contrast between the rings and gaps tends to be somewhat less in 3D than in 2D. Nevertheless, rings and gaps have clearly developed and persisted until the end of the simulation. It can also be seen from the animated version of Fig. 10 (see its caption for a link to the animation).

Refer to caption
Figure 14: Comparison of the 2D and 3D meridional structures at a representative time t=1200𝑡1200t=1200italic_t = 1200 years. The left panels plot the logarithmic scale of plasma βθsubscript𝛽𝜃\beta_{\theta}italic_β start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT (based on the polar component of the magnetic field), with velocity streamlines traced by the gray lines with arrows. The right panels plot the radial mass flux per unit area, with the magnetic field lines traced by the white lines with arrows. The top panels are for the axisymmetric model T-2D, and the lower three panels are for three meridional planes of the 3D model T-3D-0 with, respectively, ϕ=0italic-ϕ0\phi=0italic_ϕ = 0, 2π/32𝜋32\pi/32 italic_π / 3, and4π/34𝜋34\pi/34 italic_π / 3. The brown and black contour lines mark where the mass density normalized to its initial value is 0.6 and 1.0, respectively.

We believe the ring and gap formation persists because the RWI modifies but does not fundamentally change the flow structures that enable the formation and growth of rings and gaps. This is illustrated in Fig. 14. The top left panel shows a much lower plasma-βθsubscript𝛽𝜃\beta_{\theta}italic_β start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT (based on the θ𝜃\thetaitalic_θ-component of the magnetic field) in the low-density gaps than in the dense rings, consistent with the result from earlier work that the poloidal magnetic flux tends to be redistributed from the rings to the gaps. This poloidal magnetic flux redistribution pattern is fundamental to the MHD mechanism of ring and gap formation (e.g., Suriano et al., 2018; Riols & Lesur, 2019; Cui & Bai, 2021). It is broadly preserved in 3D, as seen in the three lower left panels, which plot the meridional distributions of the plasma-βθsubscript𝛽𝜃\beta_{\theta}italic_β start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT at three representative azimuthal angles. Also evident from the panels are azimuthal variations of the density (particularly in the rings between 10similar-toabsent10\sim 10∼ 10 and 15151515 au) and the plasma-βθsubscript𝛽𝜃\beta_{\theta}italic_β start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT (particularly around 16similar-toabsent16\sim 16∼ 16 au where its value is much higher at ϕ=4π/3italic-ϕ4𝜋3\phi=4\pi/3italic_ϕ = 4 italic_π / 3 than at ϕ=0italic-ϕ0\phi=0italic_ϕ = 0 and 2π/32𝜋32\pi/32 italic_π / 3). Similarly, the characteristic disk flow structure in 2D, with the accretion concentrating in warped layers (especially within about 20 au radius where most rings and gaps are located; see the blue regions in the top right panel), is broadly preserved in 3D. Suriano et al. (2018) argued that this flow structure is important for ring and gap formation because it generates sharp pinching of the poloidal magnetic field, facilitating reconnection and magnetic flux redistribution relative to mass. There are azimuthal variations in the accretion flow, as seen in the three lower-right panels around, e.g., the similar-to\sim15-au radius. They appear to modify but not destroy the rings and gaps.

5 Conclusion

We performed 2D (axisymmetric) and fully 3D non-ideal MHD simulations of substructure formation in magnetized, weakly ionized, wind-launching disks, with the non-ideal MHD coefficients computed from a simplified chemical network including dust grains. A 2D simulation using a simple power-law prescription for the ionization fraction is also included for comparison. We focused on whether the substructures formed in such disks are stable to the Rossby Wave Instability (RWI) and, if not, how the RWI affects the formation and evolution of the substructures. Our main conclusions are summarized as follows.

  1. 1.

    In agreement with previous work, axisymmetric gas rings and gaps form spontaneously in 2D simulations of non-ideal MHD disks with ambipolar diffusion and Ohmic dissipation, with more prominent substructures for larger Elsasser numbers (or better magnetic coupling). The case for non-ideal MHD formation of substructures is thus strengthened, at least under the assumption of axisymmetry.

  2. 2.

    Axisymmetric gas rings in 2D simulations are found to be unstable to Rossby Wave Instability (RWI) according to analytic stability conditions and through 3D restart simulation. As expected for RWI, shorter wavelength (or larger m𝑚mitalic_m) azimuthal modes develop earlier in the simulation, and longer wavelength ones dominate later, forming elongated anti-cyclonic vortices (arcs) in the ring’s column density distribution that last until the end of the simulation. Highly elongated vortices with aspect ratios of 10 or more are found to decay with time in our simulation, in contrast with the hydro case. This difference could be caused by magnetically induced motions, particularly strong meridional circulations with large values of azimuthal vorticity, which may be incompatible with the columnar structure preferred by vortices. Axisymmetric gaps in the 2D simulation are also unstable to RWI, forming cyclonic vortices despite being strongly magnetized, with a plasma-β𝛽\betaitalic_β of order unity. Nevertheless, the cyclonic and anti-cyclonic RWI vortices modify but do not destroy the rings and gaps in the radial gas distribution of the disk.

  3. 3.

    Our 3D simulation that starts from a smooth initial condition shows that RWI does not shut off the mechanism for generating rings and gaps in the non-ideal MHD and wind-launching disk. Specifically, anti-cyclonic and cyclonic vortices are still formed in rings and gaps, respectively, with amplitudes saturating at moderate levels. The RWI and associated vortices modify but do not suppress the poloidal magnetic flux accumulation in low-density regions and the characteristic meridional flow patterns that are crucial to the ring and gap formation.

An interesting question to address in future investigations is how RWI affects the dust distribution and the disk appearance in continuum emission. Most observed disk dust substructures appear rather axisymmetric, although a small fraction shows non-axisymmetric features, including arcs, which are potentially vortices. The RWI vortices discussed in this paper open up the possibility of producing such dusty vortices in non-ideal MHD wind-launching disks. However, detailed dust treatment is needed to explore this possibility quantitatively. The current simulation will also benefit from adaptive (rather than static) mesh refinement, which will better resolve the vortices formed in the global 3D simulations.

Acknowledgements

CYH acknowledges support from NASA SOFIA grant SOF-10_0505 and NRAO ALMA Student Observing Support (SOS) and computing resources from UVA research computing (RIVANNA) and NASA High-Performance Computing. ZYL is supported in part by NASA 80NSSC20K0533 and NSF AST-2307199. MKL is supported by the National Science and Technology Council (grants 111-2112- M-001-062-, 112-2112-M-001-064-, 111-2124-M-002-013- , 112-2124-M-002-003-) and an Academia Sinica Career Development Award (AS-CDA110-M06).

Data Availability

The inclusion of a Data Availability Statement is a requirement for articles published in MNRAS. Data Availability Statements provide a standardised format for readers to understand the availability of data underlying the research results described in the article. The statement may refer to original data generated in the course of the study or to third-party data analysed in the article. The statement should describe and provide means of access, where possible, by linking to the data or providing the required accession numbers for the relevant databases or DOIs.

References

  • ALMA Partnership et al. (2015) ALMA Partnership et al., 2015, ApJ, 808, L3
  • Andrews et al. (2018) Andrews S. M., et al., 2018, ApJ, 869, L41
  • Bae et al. (2017) Bae J., Zhu Z., Hartmann L., 2017, ApJ, 850, 201
  • Bae et al. (2023) Bae J., Isella A., Zhu Z., Martin R., Okuzumi S., Suriano S., 2023, in Inutsuka S., Aikawa Y., Muto T., Tomida K., Tamura M., eds, Astronomical Society of the Pacific Conference Series Vol. 534, Protostars and Planets VII. p. 423 (arXiv:2210.13314), doi:10.48550/arXiv.2210.13314
  • Bai & Stone (2013) Bai X.-N., Stone J. M., 2013, ApJ, 769, 76
  • Bai & Stone (2017) Bai X.-N., Stone J. M., 2017, ApJ, 836, 46
  • Boehler et al. (2021) Boehler Y., et al., 2021, A&A, 650, A59
  • Cimerman & Rafikov (2023) Cimerman N. P., Rafikov R. R., 2023, MNRAS, 519, 208
  • Cui & Bai (2021) Cui C., Bai X.-N., 2021, MNRAS, 507, 1106
  • Dong et al. (2015) Dong R., Zhu Z., Whitney B., 2015, ApJ, 809, 93
  • Draine & Sutin (1987) Draine B. T., Sutin B., 1987, ApJ, 320, 803
  • Flock et al. (2015) Flock M., Ruge J. P., Dzyurkevich N., Henning T., Klahr H., Wolf S., 2015, A&A, 574, A68
  • Galametz et al. (2018) Galametz M., et al., 2018, A&A, 616, A139
  • Grassi et al. (2019) Grassi T., Padovani M., Ramsey J. P., Galli D., Vaytet N., Ercolano B., Haugbølle T., 2019, MNRAS, 484, 161
  • Hu et al. (2022) Hu X., Li Z.-Y., Zhu Z., Yang C.-C., 2022, MNRAS, 516, 2006
  • Johansen et al. (2009) Johansen A., Youdin A., Klahr H., 2009, ApJ, 697, 1269
  • Krapp et al. (2018) Krapp L., Gressel O., Benítez-Llambay P., Downes T. P., Mohandas G., Pessah M. E., 2018, ApJ, 865, 105
  • Le Gouellec et al. (2020) Le Gouellec V. J. M., et al., 2020, A&A, 644, A11
  • Lesur (2021) Lesur G. R. J., 2021, A&A, 650, A35
  • Lesur & Papaloizou (2009) Lesur G., Papaloizou J. C. B., 2009, A&A, 498, 1
  • Lesur et al. (2022) Lesur G., et al., 2022, arXiv e-prints, p. arXiv:2203.09821
  • Li et al. (2000) Li H., Finn J. M., Lovelace R. V. E., Colgate S. A., 2000, ApJ, 533, 1023
  • Li et al. (2001) Li H., Colgate S. A., Wendroff B., Liska R., 2001, ApJ, 551, 874
  • Lin (2012) Lin M.-K., 2012, ApJ, 754, 21
  • Lin (2013) Lin M.-K., 2013, ApJ, 765, 84
  • Liu et al. (2003) Liu B., Goree J., Nosenko V., Boufendi L., 2003, Physics of Plasmas, 10, 9
  • Lovelace et al. (1999) Lovelace R. V. E., Li H., Colgate S. A., Nelson A. F., 1999, ApJ, 513, 805
  • Lyra & Klahr (2011) Lyra W., Klahr H., 2011, A&A, 527, A138
  • Lyra & Mac Low (2012) Lyra W., Mac Low M.-M., 2012, ApJ, 756, 62
  • Lyra et al. (2008) Lyra W., Johansen A., Klahr H., Piskunov N., 2008, A&A, 491, L41
  • Marr & Dong (2022) Marr M., Dong R., 2022, ApJ, 930, 80
  • Mathis et al. (1977) Mathis J. S., Rumpl W., Nordsieck K. H., 1977, ApJ, 217, 425
  • Maury et al. (2018) Maury A. J., et al., 2018, MNRAS, 477, 2760
  • McElroy et al. (2013) McElroy D., Walsh C., Markwick A. J., Cordiner M. A., Smith K., Millar T. J., 2013, A&A, 550, A36
  • Meheut et al. (2010) Meheut H., Casse F., Varniere P., Tagger M., 2010, A&A, 516, A31
  • Meheut et al. (2012) Meheut H., Yu C., Lai D., 2012, MNRAS, 422, 2399
  • Nishi et al. (1991) Nishi R., Nakano T., Umebayashi T., 1991, ApJ, 368, 181
  • Nolan et al. (2023) Nolan C. A., Zhao B., Caselli P., Li Z. Y., 2023, MNRAS, 525, 5450
  • Okuzumi et al. (2016) Okuzumi S., Momose M., Sirono S.-i., Kobayashi H., Tanaka H., 2016, ApJ, 821, 82
  • Ono et al. (2016) Ono T., Muto T., Takeuchi T., Nomura H., 2016, ApJ, 823, 84
  • Ono et al. (2018) Ono T., Muto T., Tomida K., Zhu Z., 2018, ApJ, 864, 70
  • Pinto et al. (2008) Pinto C., Galli D., Bacciotti F., 2008, A&A, 484, 1
  • Richard et al. (2013) Richard S., Barge P., Le Dizès S., 2013, A&A, 559, A30
  • Riols & Lesur (2019) Riols A., Lesur G., 2019, A&A, 625, A108
  • Ruge et al. (2016) Ruge J. P., Flock M., Wolf S., Dzyurkevich N., Fromang S., Henning T., Klahr H., Meheut H., 2016, A&A, 590, A17
  • Shu (1992) Shu F. H., 1992, Physics of Astrophysics, Vol. II
  • Stone et al. (2020) Stone J. M., Tomida K., White C. J., Felker K. G., 2020, ApJS, 249, 4
  • Suriano et al. (2017) Suriano S. S., Li Z.-Y., Krasnopolsky R., Shang H., 2017, MNRAS, 468, 3850
  • Suriano et al. (2018) Suriano S. S., Li Z.-Y., Krasnopolsky R., Shang H., 2018, MNRAS, 477, 1239
  • Takahashi & Inutsuka (2016) Takahashi S. Z., Inutsuka S.-i., 2016, AJ, 152, 184
  • Tominaga et al. (2020) Tominaga R. T., Takahashi S. Z., Inutsuka S.-i., 2020, ApJ, 900, 182
  • Umebayashi & Nakano (1990) Umebayashi T., Nakano T., 1990, MNRAS, 243, 103
  • Umurhan (2010) Umurhan O. M., 2010, A&A, 521, A25
  • Varnière & Tagger (2006) Varnière P., Tagger M., 2006, A&A, 446, L13
  • Wardle (2007) Wardle M., 2007, Ap&SS, 311, 35
  • Zanni et al. (2007) Zanni C., Ferrari A., Rosner R., Bodo G., Massaglia S., 2007, A&A, 469, 811
  • Zaqarashvili et al. (2021) Zaqarashvili T. V., et al., 2021, Space Sci. Rev., 217, 15
  • Zhao et al. (2018) Zhao B., Caselli P., Li Z.-Y., 2018, MNRAS, 478, 2723

Appendix A Charge Abundances and Elsasser Numbers

One of our improvements over previous work along a similar line (Suriano et al., 2018; Hu et al., 2022) is that we computed the charge abundances from a reduced chemical network (see § 3) rather than using the simple power-law prescription of Shu (1992), which is derived under the assumption of one dominant molecular ion (e.g., HCO+superscriptHCO{\rm HCO^{+}}roman_HCO start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT) and no dust grains. Fig. 15 shows fractional abundances of the main charges computed from our chemical network as a function of number density for an MRN-like dust size distribution with amin=0.5μsubscript𝑎min0.5𝜇a_{\rm min}=0.5\muitalic_a start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT = 0.5 italic_μm and amax=25μsubscript𝑎max25𝜇a_{\rm max}=25\muitalic_a start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 25 italic_μm, including Mg+superscriptMg{\rm Mg^{+}}roman_Mg start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT, m+superscriptm{\rm m^{+}}roman_m start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT (molecular ions other than H3+superscriptsubscriptH3{\rm H_{3}^{+}}roman_H start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT), grains with charge 0,±e0plus-or-minus𝑒0,\pm e0 , ± italic_e, and ±2eplus-or-minus2𝑒\pm 2e± 2 italic_e. The power-law prescription of Shu (1992) is also shown for comparison.

We use the charge abundances in Fig. 15 to calculate the Elsasser number together with the disk properties. Fig. 16 shows the initial profile of the ambipolar Elsasser numbers of Model S-2D (where the power-law prescription of Shu (1992) is adopted; green lines) and Model T-2D (where the chemical network is used; red lines) as a function of radius on the mid-plane (panel a) and as a function of vertical distance from the midplane at r=10𝑟10r=10italic_r = 10 au (panel b, where the θ𝜃\thetaitalic_θ-dependence in equation [25] is applied).

Refer to caption
Figure 15: Fractional abundances of main charges as a function of the number density computed from the chemical network for an MRN-type dust size distribution with amin=0.5μsubscript𝑎min0.5𝜇a_{\rm min}=0.5\muitalic_a start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT = 0.5 italic_μm and amax=25μsubscript𝑎max25𝜇a_{\rm max}=25\muitalic_a start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 25 italic_μm. Magnesium ions, molecular ions other than H3+superscriptsubscriptH3{\rm H_{3}^{+}}roman_H start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT, grains with charge 0,±e0plus-or-minus𝑒0,\pm e0 , ± italic_e, and ±2eplus-or-minus2𝑒\pm 2e± 2 italic_e. The power-law prescription of Shu (1992) is plotted as a yellow line for comparison.
Refer to caption
Figure 16: Initial profiles of the ambipolar Elsasser number for Model S-2D (green lines) and Model T-2D (red lines). Panel (a) shows the Elsasser numbers as a function of radius on the mid-plane. Panel (b) shows the Elsasser number at the radius r=10𝑟10r=10italic_r = 10 au as a function of the vertical distance from the mid-plane. The dashed lines in panel (b) mark the disc surface near similar-to\sim 1 au.
Table 2: Relative abundances of the elements included in the chemical network. Similar to Table 1 of Umebayashi & Nakano (1990), with the fractions remaining in the gas phase set to δ1=0.2subscript𝛿10.2\delta_{1}=0.2italic_δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 0.2 and δ2=0.02subscript𝛿20.02\delta_{2}=0.02italic_δ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 0.02.
Element Abundant Chemical species Fraction (gas)
HH{\rm H}roman_H 1 H2subscriptH2{\rm H_{2}}roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT 1
HeHe{\rm He}roman_He 8.5 ×102absentsuperscript102\times 10^{-2}× 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT HeHe{\rm He}roman_He 1
CC{\rm C}roman_C 4.2 ×104absentsuperscript104\times 10^{-4}× 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT COCO{\rm CO}roman_CO δ1subscript𝛿1\delta_{1}italic_δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT
OO{\rm O}roman_O 6.9 ×104absentsuperscript104\times 10^{-4}× 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT
9.0 ×105absentsuperscript105\times 10^{-5}× 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT OO{\rm O}roman_O δ1subscript𝛿1\delta_{1}italic_δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT
9.0 ×105absentsuperscript105\times 10^{-5}× 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT O2subscriptO2{\rm O_{2}}roman_O start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT δ1subscript𝛿1\delta_{1}italic_δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT
MgMg{\rm Mg}roman_Mg 1.7 ×104absentsuperscript104\times 10^{-4}× 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT MgMg{\rm Mg}roman_Mg δ2subscript𝛿2\delta_{2}italic_δ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT
Table 3: List of gas-phase reactions in our chemical network. Symbols αcsubscript𝛼c\alpha_{\rm c}italic_α start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT, βcsubscript𝛽c\beta_{\rm c}italic_β start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT, and γcsubscript𝛾c\gamma_{\rm c}italic_γ start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT are the coefficients used to determine the reaction rates. These coefficients are grabbed from UMIST database.
Reaction αcsubscript𝛼c\alpha_{\rm c}italic_α start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT βcsubscript𝛽c\beta_{\rm c}italic_β start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT γcsubscript𝛾c\gamma_{\rm c}italic_γ start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT
H++OO++HsuperscriptHOsuperscriptOH{\rm H}^{+}+{\rm O}\longrightarrow{\rm O}^{+}+{\rm H}roman_H start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_O ⟶ roman_O start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_H 6.86×10106.86superscript10106.86\times 10^{-10}6.86 × 10 start_POSTSUPERSCRIPT - 10 end_POSTSUPERSCRIPT 0.26 224.30
H++O2O2++HsuperscriptHsubscriptO2superscriptsubscriptO2H{\rm H}^{+}+{\rm O_{2}}\longrightarrow{\rm O_{2}}^{+}+{\rm H}roman_H start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_O start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟶ roman_O start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_H 2.00×1092.00superscript1092.00\times 10^{-9}2.00 × 10 start_POSTSUPERSCRIPT - 9 end_POSTSUPERSCRIPT 0.0 0.0
H++MgMg++HsuperscriptHMgsuperscriptMgH{\rm H}^{+}+{\rm Mg}\longrightarrow{\rm Mg}^{+}+{\rm H}roman_H start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_Mg ⟶ roman_Mg start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_H 1.1×1091.1superscript1091.1\times 10^{-9}1.1 × 10 start_POSTSUPERSCRIPT - 9 end_POSTSUPERSCRIPT 0.0 0.0
He++H2H++H+HesuperscriptHesubscriptH2superscriptHHHe{\rm He}^{+}+{\rm H_{2}}\longrightarrow{\rm H}^{+}+{\rm H}+{\rm He}roman_He start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟶ roman_H start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_H + roman_He 7.2×10157.2superscript10157.2\times 10^{-15}7.2 × 10 start_POSTSUPERSCRIPT - 15 end_POSTSUPERSCRIPT 0.0 0.0
He++COC++O+HesuperscriptHeCOsuperscriptCOHe{\rm He}^{+}+{\rm CO}\longrightarrow{\rm C}^{+}+{\rm O}+{\rm He}roman_He start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_CO ⟶ roman_C start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_O + roman_He 1.6×1091.6superscript1091.6\times 10^{-9}1.6 × 10 start_POSTSUPERSCRIPT - 9 end_POSTSUPERSCRIPT 0.0 0.0
He++O2O++O+HesuperscriptHesubscriptO2superscriptOOHe{\rm He}^{+}+{\rm O_{2}}\longrightarrow{\rm O}^{+}+{\rm O}+{\rm He}roman_He start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_O start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟶ roman_O start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_O + roman_He 1.1×1091.1superscript1091.1\times 10^{-9}1.1 × 10 start_POSTSUPERSCRIPT - 9 end_POSTSUPERSCRIPT 0.0 0.0
H3++COHCO++H2superscriptsubscriptH3COsuperscriptHCOsubscriptH2{\rm H_{3}}^{+}+{\rm CO}\longrightarrow{\rm HCO}^{+}+{\rm H_{2}}roman_H start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_CO ⟶ roman_HCO start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT 1.36×1091.36superscript1091.36\times 10^{-9}1.36 × 10 start_POSTSUPERSCRIPT - 9 end_POSTSUPERSCRIPT -0.14 -3.40
H3++OOH++H2superscriptsubscriptH3OsuperscriptOHsubscriptH2{\rm H_{3}}^{+}+{\rm O}\longrightarrow{\rm OH}^{+}+{\rm H_{2}}roman_H start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_O ⟶ roman_OH start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT 7.98×10107.98superscript10107.98\times 10^{-10}7.98 × 10 start_POSTSUPERSCRIPT - 10 end_POSTSUPERSCRIPT -0.16 1.4
H3++O2O2H++H2superscriptsubscriptH3subscriptO2subscriptO2superscriptHsubscriptH2{\rm H_{3}}^{+}+{\rm O_{2}}\longrightarrow{\rm O_{2}H}^{+}+{\rm H_{2}}roman_H start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_O start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟶ roman_O start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT roman_H start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT 9.3×10109.3superscript10109.3\times 10^{-10}9.3 × 10 start_POSTSUPERSCRIPT - 10 end_POSTSUPERSCRIPT 0.0 100.0
H3++MgMg++H2+HsuperscriptsubscriptH3MgsuperscriptMgsubscriptH2H{\rm H_{3}}^{+}+{\rm Mg}\longrightarrow{\rm Mg}^{+}+{\rm H_{2}}+{\rm H}roman_H start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_Mg ⟶ roman_Mg start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + roman_H 1.0×1091.0superscript1091.0\times 10^{-9}1.0 × 10 start_POSTSUPERSCRIPT - 9 end_POSTSUPERSCRIPT 0.0 0.0
C++H2CH2++hνsuperscriptCsubscriptH2superscriptsubscriptCH2𝜈{\rm C}^{+}+{\rm H_{2}}\longrightarrow{\rm CH_{2}}^{+}+h\nuroman_C start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟶ roman_CH start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + italic_h italic_ν 2.0×10162.0superscript10162.0\times 10^{-16}2.0 × 10 start_POSTSUPERSCRIPT - 16 end_POSTSUPERSCRIPT -1.3 23.0
C++O2CO++OsuperscriptCsubscriptO2superscriptCOO{\rm C}^{+}+{\rm O_{2}}\longrightarrow{\rm CO}^{+}+{\rm O}roman_C start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_O start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟶ roman_CO start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_O 3.42×10103.42superscript10103.42\times 10^{-10}3.42 × 10 start_POSTSUPERSCRIPT - 10 end_POSTSUPERSCRIPT 0.0 0.0
C++O2O++COsuperscriptCsubscriptO2superscriptOCO{\rm C}^{+}+{\rm O_{2}}\longrightarrow{\rm O}^{+}+{\rm CO}roman_C start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_O start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟶ roman_O start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_CO 4.54×10104.54superscript10104.54\times 10^{-10}4.54 × 10 start_POSTSUPERSCRIPT - 10 end_POSTSUPERSCRIPT 0.0 0.0
C++MgMg++CsuperscriptCMgsuperscriptMgC{\rm C}^{+}+{\rm Mg}\longrightarrow{\rm Mg}^{+}+{\rm C}roman_C start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_Mg ⟶ roman_Mg start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_C 1.1×1091.1superscript1091.1\times 10^{-9}1.1 × 10 start_POSTSUPERSCRIPT - 9 end_POSTSUPERSCRIPT 0.0 0.0
O2++MgMg++O2superscriptsubscriptO2MgsuperscriptMgsubscriptO2{\rm O_{2}}^{+}+{\rm Mg}\longrightarrow{\rm Mg}^{+}+{\rm O_{2}}roman_O start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_Mg ⟶ roman_Mg start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_O start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT 1.2×1091.2superscript1091.2\times 10^{-9}1.2 × 10 start_POSTSUPERSCRIPT - 9 end_POSTSUPERSCRIPT 0.0 0.0
HCO++MgMg++HCOsuperscriptHCOMgsuperscriptMgHCO{\rm HCO}^{+}+{\rm Mg}\longrightarrow{\rm Mg}^{+}+{\rm HCO}roman_HCO start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_Mg ⟶ roman_Mg start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_HCO 2.9×1092.9superscript1092.9\times 10^{-9}2.9 × 10 start_POSTSUPERSCRIPT - 9 end_POSTSUPERSCRIPT 0.0 0.0
H++eH+hνsuperscriptHeH𝜈{\rm H}^{+}+{\rm e}\longrightarrow{\rm H}+h\nuroman_H start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_e ⟶ roman_H + italic_h italic_ν 3.5×10123.5superscript10123.5\times 10^{-12}3.5 × 10 start_POSTSUPERSCRIPT - 12 end_POSTSUPERSCRIPT -0.75 0.0
He++eHe+hνsuperscriptHeeHe𝜈{\rm He}^{+}+{\rm e}\longrightarrow{\rm He}+h\nuroman_He start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_e ⟶ roman_He + italic_h italic_ν 5.36×10125.36superscript10125.36\times 10^{-12}5.36 × 10 start_POSTSUPERSCRIPT - 12 end_POSTSUPERSCRIPT -0.50 0.0
H3++eH+H+HsuperscriptsubscriptH3eHHH{\rm H_{3}}^{+}+{\rm e}\longrightarrow{\rm H}+{\rm H}+{\rm H}roman_H start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_e ⟶ roman_H + roman_H + roman_H 4.36×1084.36superscript1084.36\times 10^{-8}4.36 × 10 start_POSTSUPERSCRIPT - 8 end_POSTSUPERSCRIPT -0.52 0.0
H3++eH2+HsuperscriptsubscriptH3esubscriptH2H{\rm H_{3}}^{+}+{\rm e}\longrightarrow{\rm H_{2}}+{\rm H}roman_H start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_e ⟶ roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + roman_H 2.34×1082.34superscript1082.34\times 10^{-8}2.34 × 10 start_POSTSUPERSCRIPT - 8 end_POSTSUPERSCRIPT -0.52 0.0
C++eC+hνsuperscriptCeC𝜈{\rm C}^{+}+{\rm e}\longrightarrow{\rm C}+h\nuroman_C start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_e ⟶ roman_C + italic_h italic_ν 2.36×10122.36superscript10122.36\times 10^{-12}2.36 × 10 start_POSTSUPERSCRIPT - 12 end_POSTSUPERSCRIPT -0.29 -17.6
Mg++eMg+hνsuperscriptMgeMg𝜈{\rm Mg}^{+}+{\rm e}\longrightarrow{\rm Mg}+h\nuroman_Mg start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_e ⟶ roman_Mg + italic_h italic_ν 2.78×10122.78superscript10122.78\times 10^{-12}2.78 × 10 start_POSTSUPERSCRIPT - 12 end_POSTSUPERSCRIPT -0.68 0.0
CO++eO+CsuperscriptCOeOC{\rm CO}^{+}+{\rm e}\longrightarrow{\rm O}+{\rm C}roman_CO start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_e ⟶ roman_O + roman_C 2.0×1072.0superscript1072.0\times 10^{-7}2.0 × 10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT -0.48 0.0
O2++eO+OsuperscriptsubscriptO2eOO{\rm O_{2}}^{+}+{\rm e}\longrightarrow{\rm O}+{\rm O}roman_O start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_e ⟶ roman_O + roman_O 1.95×1071.95superscript1071.95\times 10^{-7}1.95 × 10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT -0.70 0.0
OH++eO+HsuperscriptOHeOH{\rm OH}^{+}+{\rm e}\longrightarrow{\rm O}+{\rm H}roman_OH start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_e ⟶ roman_O + roman_H 3.75×1083.75superscript1083.75\times 10^{-8}3.75 × 10 start_POSTSUPERSCRIPT - 8 end_POSTSUPERSCRIPT -0.50 0.0
O2H++eO2+HsubscriptO2superscriptHesubscriptO2H{\rm O_{2}H}^{+}+{\rm e}\longrightarrow{\rm O_{2}}+{\rm H}roman_O start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT roman_H start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_e ⟶ roman_O start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + roman_H 3.0×1073.0superscript1073.0\times 10^{-7}3.0 × 10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT -0.50 0.0
HCO++eCO+HsuperscriptHCOeCOH{\rm HCO}^{+}+{\rm e}\longrightarrow{\rm CO}+{\rm H}roman_HCO start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_e ⟶ roman_CO + roman_H 2.4×1072.4superscript1072.4\times 10^{-7}2.4 × 10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT -0.69 0.0
CH2++eC+H2superscriptsubscriptCH2eCsubscriptH2{\rm CH_{2}}^{+}+{\rm e}\longrightarrow{\rm C}+{\rm H_{2}}roman_CH start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + roman_e ⟶ roman_C + roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT 7.68×1087.68superscript1087.68\times 10^{-8}7.68 × 10 start_POSTSUPERSCRIPT - 8 end_POSTSUPERSCRIPT -0.60 0.0

Appendix B Rate Coefficients for Collisional Momentum Transfer

Three charged species (ions, electrons, and grains) collide with the molecular hydrogen in our model. We take the collisional momentum transfer rate coefficients from Pinto et al. (2008) and Grassi et al. (2019). Note that the units of Re,nsubscript𝑅𝑒𝑛R_{e,n}italic_R start_POSTSUBSCRIPT italic_e , italic_n end_POSTSUBSCRIPT, Ri,nsubscript𝑅𝑖𝑛R_{i,n}italic_R start_POSTSUBSCRIPT italic_i , italic_n end_POSTSUBSCRIPT, and Rg,nsubscript𝑅𝑔𝑛R_{g,n}italic_R start_POSTSUBSCRIPT italic_g , italic_n end_POSTSUBSCRIPT in this subsection are cm3s1superscriptcm3superscripts1{\rm cm^{3}s^{-1}}roman_cm start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, corresponding to the Rj,n(T)subscript𝑅𝑗𝑛𝑇R_{j,n}(T)italic_R start_POSTSUBSCRIPT italic_j , italic_n end_POSTSUBSCRIPT ( italic_T ) in the Eq. (21), so the Hall parameter is dimensionless.

For the collisional rate between electrons and molecular hydrogen, we use the rate

Re,n(T)=109T[0.535+0.203log10(T)0.163[log10(T)]2].subscript𝑅𝑒𝑛𝑇superscript109𝑇delimited-[]0.5350.203subscriptlog10T0.163superscriptdelimited-[]subscriptlog10T2\displaystyle R_{e,n}(T)=10^{-9}\sqrt{T}\left[0.535+0.203{\rm log_{10}(T)}-0.1% 63[{\rm log_{10}(T)}]^{2}\right].italic_R start_POSTSUBSCRIPT italic_e , italic_n end_POSTSUBSCRIPT ( italic_T ) = 10 start_POSTSUPERSCRIPT - 9 end_POSTSUPERSCRIPT square-root start_ARG italic_T end_ARG [ 0.535 + 0.203 roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ( roman_T ) - 0.163 [ roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ( roman_T ) ] start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] . (28)

The collisional rate between ions and molecular hydrogen is given as

Ri,n(T)=2.21παpole2mred.subscript𝑅𝑖𝑛𝑇2.21𝜋subscript𝛼polsuperscript𝑒2subscript𝑚red\displaystyle R_{i,n}(T)=2.21\pi\sqrt{\frac{\alpha_{\rm pol}e^{2}}{m_{\rm red}% }}.italic_R start_POSTSUBSCRIPT italic_i , italic_n end_POSTSUBSCRIPT ( italic_T ) = 2.21 italic_π square-root start_ARG divide start_ARG italic_α start_POSTSUBSCRIPT roman_pol end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_m start_POSTSUBSCRIPT roman_red end_POSTSUBSCRIPT end_ARG end_ARG . (29)

where αpol=8.06×1025cm3subscript𝛼pol8.06superscript1025superscriptcm3\alpha_{\rm pol}=8.06\times 10^{-25}{\rm cm^{3}}italic_α start_POSTSUBSCRIPT roman_pol end_POSTSUBSCRIPT = 8.06 × 10 start_POSTSUPERSCRIPT - 25 end_POSTSUPERSCRIPT roman_cm start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT is the polarizability of molecular hydrogen.

The momentum transfer through collisions between charged grains and neutral hydrogen is

Rg,n(T,Z)=2.21παpol|Z|e2mredacp+1aminp+1amaxp+1aminp+1delimited-⟨⟩subscript𝑅𝑔𝑛𝑇𝑍2.21𝜋subscript𝛼pol𝑍superscript𝑒2subscript𝑚redsubscriptsuperscript𝑎𝑝1csubscriptsuperscript𝑎𝑝1minsubscriptsuperscript𝑎𝑝1maxsubscriptsuperscript𝑎𝑝1min\displaystyle\langle R_{g,n}(T,Z)\rangle=2.21\pi\sqrt{\frac{\alpha_{\rm pol}|Z% |e^{2}}{m_{\rm red}}}\frac{a^{p+1}_{\rm c}-a^{p+1}_{\rm min}}{a^{p+1}_{\rm max% }-a^{p+1}_{\rm min}}⟨ italic_R start_POSTSUBSCRIPT italic_g , italic_n end_POSTSUBSCRIPT ( italic_T , italic_Z ) ⟩ = 2.21 italic_π square-root start_ARG divide start_ARG italic_α start_POSTSUBSCRIPT roman_pol end_POSTSUBSCRIPT | italic_Z | italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_m start_POSTSUBSCRIPT roman_red end_POSTSUBSCRIPT end_ARG end_ARG divide start_ARG italic_a start_POSTSUPERSCRIPT italic_p + 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT - italic_a start_POSTSUPERSCRIPT italic_p + 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT end_ARG start_ARG italic_a start_POSTSUPERSCRIPT italic_p + 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT - italic_a start_POSTSUPERSCRIPT italic_p + 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT end_ARG
+(8kBTπmr)1/24πδL3amaxp+3acp+3amaxp+1aminp+1p+1p+3,superscript8subscript𝑘B𝑇𝜋subscript𝑚r124𝜋subscript𝛿L3subscriptsuperscript𝑎𝑝3maxsubscriptsuperscript𝑎𝑝3csubscriptsuperscript𝑎𝑝1maxsubscriptsuperscript𝑎𝑝1min𝑝1𝑝3\displaystyle+\left(\frac{8k_{\rm B}T}{\pi m_{\rm r}}\right)^{1/2}\frac{4\pi% \delta_{\rm L}}{3}\frac{a^{p+3}_{\rm max}-a^{p+3}_{\rm c}}{a^{p+1}_{\rm max}-a% ^{p+1}_{\rm min}}\frac{p+1}{p+3},+ ( divide start_ARG 8 italic_k start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT italic_T end_ARG start_ARG italic_π italic_m start_POSTSUBSCRIPT roman_r end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT divide start_ARG 4 italic_π italic_δ start_POSTSUBSCRIPT roman_L end_POSTSUBSCRIPT end_ARG start_ARG 3 end_ARG divide start_ARG italic_a start_POSTSUPERSCRIPT italic_p + 3 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT - italic_a start_POSTSUPERSCRIPT italic_p + 3 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT end_ARG start_ARG italic_a start_POSTSUPERSCRIPT italic_p + 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT - italic_a start_POSTSUPERSCRIPT italic_p + 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT end_ARG divide start_ARG italic_p + 1 end_ARG start_ARG italic_p + 3 end_ARG , (30)

where

ac(T,Z)=0.206δL(αpol|Z|T)1/4subscript𝑎𝑐𝑇𝑍0.206subscript𝛿Lsuperscriptsubscript𝛼pol𝑍𝑇14\displaystyle a_{c}(T,Z)=\frac{0.206}{\sqrt{\delta_{\rm L}}}\left(\frac{\alpha% _{\rm pol}|Z|}{T}\right)^{1/4}italic_a start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_T , italic_Z ) = divide start_ARG 0.206 end_ARG start_ARG square-root start_ARG italic_δ start_POSTSUBSCRIPT roman_L end_POSTSUBSCRIPT end_ARG end_ARG ( divide start_ARG italic_α start_POSTSUBSCRIPT roman_pol end_POSTSUBSCRIPT | italic_Z | end_ARG start_ARG italic_T end_ARG ) start_POSTSUPERSCRIPT 1 / 4 end_POSTSUPERSCRIPT (31)

is a critical grain size to measure microscope grain-neutral reaction for the hard sphere and the Langevin rates (Pinto et al., 2008). We set δL=1.3subscript𝛿L1.3\delta_{\rm L}=1.3italic_δ start_POSTSUBSCRIPT roman_L end_POSTSUBSCRIPT = 1.3 as recommended by Liu et al. (2003).