11institutetext: Department of Space, Earth and Environment, Chalmers University of Technology, Gothenburg SE-412 96, Sweden 22institutetext: Department of Astronomy, University of Virginia, 530 McCormick Road, Charlottesville, VA 22904, USA 33institutetext: Department of Astronomy, San Diego State University, San Diego, CA 92182, USA 44institutetext: Computational Science Research Center, San Diego State University, San Diego, CA 92182, USA 55institutetext: Faculty of Physics, University of Duisburg-Essen, Lotharstraße 1, 47057, Duisburg, Germany

The High-resolution Accretion Disks of Embedded protoStars (HADES) simulations. I. Impact of Protostellar Magnetic Fields on the Accretion Modes

Brandt A. L. Gaches E-mail: brandt.gaches@chalmers.se11    Jonathan C. Tan , 1122    Anna L. Rosen , 3344    Rolf Kuiper 55
(Accepted XXX. Received YYY; in original form ZZZ)

How embedded, actively accreting low-mass protostars accrete their mass is still greatly debated. Observations are now piecing together the puzzle of embedded protostellar accretion, in particular with new facilities in the near-infrared. However, high-resolution theoretical models are still lacking, with a stark paucity of detailed simulations of these early phases. Here we present high-resolution non-ideal magneto-hydrodynamic simulations of a Solar mass protostar accreting at rates exceeding 10M6superscriptsubscript𝑀direct-product6{}^{-6}M_{\odot}start_FLOATSUPERSCRIPT - 6 end_FLOATSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT yr-1. We show the results of the accretion flow for four different protostellar magnetic fields, 10 G, 500 G, 1 kG, and 2 kG, combined with a disk magnetic field. For weaker (10 G and 500 G) protostar magnetic fields, accretion occurs via a turbulent boundary layer mode, with disk material impacting across the protostellar surface. In the 500 G model, the presence of a magnetically dominated outflow focuses the accretion towards the equator, slightly enhancing and ordering the accretion. For kG magnetic fields, the disk becomes truncated due to the protostellar dipole and exhibits magnetospheric accretion, with the 2 kG model having accretion bursts induced by the interchange instability. We present bolometric light curves for the models and find that they reproduce observations of Class I protostars from YSOVAR, with high bursts followed by an exponential decay possibly being a signature of instability-driven accretion. Finally, we present the filling fractions of accretion and find that 90% of the mass is accreted in a surface area fraction of 10-20%. These simulations will be extended in future work for a broader parameter space, with their high resolution and high temporal spacing able to explore a wide range of interesting protostellar physics.

Key Words.:
Stars: protostars – Accretion, accretion disks – Methods: numerical

1 Introduction

Protostars are the bloated embryos of main-sequence stars that are actively accreting material from their natal environments. In the earlier phases of evolution, low-mass protostars are still deeply embedded within their natal cores and surrounded by relatively massive disks, resulting in substantial accretion rates, m˙107Msubscript˙𝑚superscript107subscript𝑀direct-product\dot{m}_{*}\geq 10^{-7}M_{\odot}over˙ start_ARG italic_m end_ARG start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT ≥ 10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT yr-1. Direct knowledge of how protostars accrete their mass typically comes from observations at more evolved stages, such as T-Tauri stars, when the central protostar is exposed, the disks have been largely depleted and the accretion rates drop below m˙108Msubscript˙𝑚superscript108subscript𝑀direct-product\dot{m}_{*}\leq 10^{-8}M_{\odot}over˙ start_ARG italic_m end_ARG start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT ≤ 10 start_POSTSUPERSCRIPT - 8 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT yr-1. Since these objects are exposed, it becomes significantly more feasible to observe emission from the accretion zones, primarily in the form of X-rays (Feigelson & Montmerle, 1999; Feigelson et al., 2007), ultraviolet emission (Manara et al., 2023) and hydrogen recombination emission (Fischer et al., 2023). There are indirect methods to estimate the accretion rate, such as determining the properties of protostellar outflows and inferring the accretion rate via theoretical models (e.g. Avison et al., 2021). Modern observational facilities, such as the Very Large Telescope Interferometer (VLTI) with the GRAVITY instrument, are now probing the accretion of more embedded systems, tracing evolutionary times when most of the protostar’s mass has already been accumulated (e.g., Gravity Collaboration et al., 2023b, a). During the main accretion phases of protostar formation, the protostar’s natal core and disk obscure the accretion zone, complicating efforts to directly probe the inner accretion disks and follow how the material is delivered onto the protostar.

Multi-physics star formation simulations can capture the earliest phases of star formation from the large scales (sub-parsec for core-scale simulations and 1-10s of parsec for molecular cloud simulations) (e.g. Dale et al., 2012; Rosen et al., 2016; Offner & Chaban, 2017; Seifried et al., 2017; Cunningham et al., 2018; Wurster et al., 2019; Rosen & Krumholz, 2020; Grudić et al., 2021; Rosen, 2022; Suin et al., 2024), but must use sub-grid prescriptions to model the protostar as an unmagnetized sink particle which accretes mass from an assumed accretion region (Offner et al., 2009; Bleuler & Teyssier, 2014). These simulations are broadly able to reproduce the accretion rates from analytic accretion models (i.e. Shu, 1977; Bonnell et al., 2001; McKee & Tan, 2003). However, these parsec-scale simulations cannot resolve the accretion flow directly onto the protostar due to resolution constraints: resolving from molecular cloud scales to the protostellar surface requires spanning 7-8 orders of magnitude across length scales. The combined theoretical and observational challenges mean the underlying physical processes of protostellar accretion have been computationally and observationally unfeasible to probe during the main growth phases.

There have been several different regimes of protostellar accretion proposed. Boundary layer accretion (Kley & Lin, 1996) occurs when the surrounding disk reaches the protostellar surface, directly feeding material onto the protostar. Magnetospheric accretion (Koenigl, 1991; Matt & Pudritz, 2005; Rosen et al., 2012; Hartmann et al., 2016) occurs when the magnetic pressure associated with the protostar’s surface magnetic field truncates the disk, balancing the ram pressure of the accretion flow, funneling the gas along magnetic flux tubes from the disk to the protostellar surface. Analytic expressions for the radius at which the disk is truncated typically assume the field consists solely of a dipole, although the magnetic field of evolved protostars is found to be a complex mixture of multipoles (Donati et al., 2007, 2008; Gregory et al., 2008; Carroll et al., 2012). The flow from the truncated disk is still likely dominated by the dipole component, with the higher order poles perturbing the flow closer to the surface (Lovelace et al., 2010). These filaments impact the protostar and create a shock near the surface, shock-heating up to millions of degrees Kelvin causing the accretion shock to primarily radiate X-rays via thermal Bremsstrahlung cooling (Calvet & Gullbring, 1998; Hartmann et al., 2016). Accretion can also occur via funnel flows along the outflow cavity wall and through failed winds (Takasao et al., 2018) which occur when outflowing gas fails to escape the protostar’s gravitational potential. Finally, instability-driven accretion can occur via the development of fluid instabilities such as the magnetic Rayleigh-Taylor instability (also called the interchange instability) at the truncated disk surface (e.g. Kulkarni & Romanova, 2008; Zhu et al., 2024). For actively accreting, embedded protostars, the underlying mechanisms remain largely unknown, motivating the need for high-resolution multi-dimensional numerical simlations of accretion disks and their host protostars to resolve the accretion flow onto the protostar.

High-resolution simulations of protostellar accretion have primarily focused on later evolutionary times, in particular the T-Tauri phase when the protostar is exposed, primarily motivated by the availability of observational data (Zanni & Ferreira, 2009; Romanova et al., 2011; Yuan et al., 2012; Zanni & Ferreira, 2013; Blinova et al., 2016; Takasao et al., 2018; Ireland et al., 2021; Romanova et al., 2021; Takasao et al., 2022; Zhu et al., 2024). In these simulations, the disk mass is typically less than 1% of the protostar’s mass, resulting in a thin, weakly turbulent, viscous disk. The resulting accretion rates range between m˙1011108subscript˙𝑚superscript1011superscript108\dot{m}_{*}\approx 10^{-11}-10^{-8}over˙ start_ARG italic_m end_ARG start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT ≈ 10 start_POSTSUPERSCRIPT - 11 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT - 8 end_POSTSUPERSCRIPT M yr-1. Conversely, at the earliest evolutionary times, high-resolution multi-physics simulations have been performed of the collapse of a protostar core to just after its initial formation, only following the first and second collapse phases to stellar densities (Tomida et al., 2013; Bate et al., 2014; Bhandare et al., 2020; Ahmad et al., 2024). There is, thus, a significant gap in the simulations of embedded protostars during the main growth phase, when the protostar is actively accreting the majority of its stellar mass at rates of m˙107Mgreater-than-or-equivalent-tosubscript˙𝑚superscript107subscript𝑀direct-product\dot{m}_{*}\gtrsim 10^{-7}M_{\odot}over˙ start_ARG italic_m end_ARG start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT ≳ 10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT yr-1, which resolve the underlying accretion flow onto the protostellar surface.

To address this gap, we perform a suite of high-resolution non-ideal viscous magneto-hydrodynamic simulations (MHD) of accretion onto a solar mass protostar to bridge this important gap. In order to determine the importance of how the protostar’s magnetic field affects the accretion flow structure and delivery of material onto the protostellar surface, we vary the protostar’s magnetic field across a wide physical range and include non-ideal effects. Our simulations include Ohmic dissipation and the Hall effect, but we neglect ambipolar diffusion in this study. While ambipolar diffusion can be important on the larger disk and protostellar core scales (Lesur et al., 2014; Commerçon et al., 2022), the Hall effect likely dominates in the inner accretion disk (Lesur, 2021). The role of ambipolar diffusion in the inner hot accretion disk around protostars, in particular because the inner regions are more strongly ionized than the rest of the disk and core, has not been well investigated, compared to the larger scales where the gas is colder and more neutral.

Our simulations presented herein focus on a 1 M bloated protostar analog with accretion rates (m˙107˙𝑚superscript107\dot{m}\geq 10^{-7}over˙ start_ARG italic_m end_ARG ≥ 10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT M yr-1) coinciding with recent observational surveys of Fiorellino et al. (2022, 2023) of young, accreting protostars. Since theoretical studies have shown that the protostellar magnetic field plays a major role in the underlying accretion mechanism in evolved protostellar objects (see references above), we focus primarily on varying the protostellar magnetic field from 10 G to 2 kG and detailing the resulting accretion modes. Our Paper II, Chowdhury et al. (in prep.), will detail more quantitatively the outflow physics. In Section 2 we describe the physics included in the simulations and the initial and boundary conditions imposed. In Section 3 we present our results of the simulations and describe the accretion flow onto the protostar and the interplay between the protostar’s accretion rate and and bolometric luminosity variability. In Section 4 we discuss the results of these numerical experiments in the context of currently known protostar physics. Finally, we conclude and summarize our results in Section 5.

2 Methods

We perform four simulations amounting to numerical experiments to investigate the role of the protostellar magnetic field on the underlying accretion physics from its surrounding protostellar disk onto the protostellar surface. We use the public Pluto MHD simulation code, which includes non-ideal MHD terms for Ohmic resistivity and the Hall effect, shear viscosity, and an external gravitational potential. We neglect ambipolar diffusion in this study. As stated above, we neglect ambipolar diffusion in our simulations because it is likely subdominant to the other non-ideal effects for the innermost regions of the protostellar accretion disk. We use the HLL solver with the Lagrange multiplier divergence cleaning method, to enforce B=0@vecB0\nabla\cdot\@vec{B}=0∇ ⋅ start_ID start_ARG italic_B end_ARG end_ID = 0, and the van Leer limiter. Pressure is updated based on the entropy of the gas, which is advected as a passive scalar. The equations are evolved using a second-order Runge-Kutta integration scheme. The equations solved by Pluto are

ρt+(ρv)=0𝜌𝑡𝜌@vecv0\frac{\partial\rho}{\partial t}+\nabla\cdot(\rho\@vec{v})=0divide start_ARG ∂ italic_ρ end_ARG start_ARG ∂ italic_t end_ARG + ∇ ⋅ ( italic_ρ start_ID start_ARG italic_v end_ARG end_ID ) = 0 (1)
(ρv)t+[ρvvBB+𝕀(p+B22)]T=ρΦ+Π𝜌@vecv𝑡superscriptdelimited-[]𝜌@vecv@vecv@vecB@vecB𝕀𝑝superscript𝐵22𝑇𝜌ΦΠ\frac{\partial(\rho\@vec{v})}{\partial t}+\nabla\cdot\left[\rho\@vec{v}\@vec{v% }-\@vec{B}\@vec{B}+\mathbb{I}\left(p+\frac{B^{2}}{2}\right)\right]^{T}=-\rho% \nabla\Phi+\nabla\cdot\Pidivide start_ARG ∂ ( italic_ρ start_ID start_ARG italic_v end_ARG end_ID ) end_ARG start_ARG ∂ italic_t end_ARG + ∇ ⋅ [ italic_ρ start_ID start_ARG italic_v end_ARG end_ID start_ID start_ARG italic_v end_ARG end_ID - start_ID start_ARG italic_B end_ARG end_ID start_ID start_ARG italic_B end_ARG end_ID + blackboard_I ( italic_p + divide start_ARG italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG ) ] start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT = - italic_ρ ∇ roman_Φ + ∇ ⋅ roman_Π (2)
Bt+×(cE)=0@vecB𝑡𝑐@vecE0\frac{\partial\@vec{B}}{\partial t}+\nabla\times\left(c\@vec{E}\right)=0divide start_ARG ∂ start_ID start_ARG italic_B end_ARG end_ID end_ARG start_ARG ∂ italic_t end_ARG + ∇ × ( italic_c start_ID start_ARG italic_E end_ARG end_ID ) = 0 (3)
(Et+ρΦ)t+[(ρv22+ρe+p+ρΦ)v+cE×B]=(vΠ),subscript𝐸𝑡𝜌Φ𝑡delimited-[]𝜌superscript@vecv22𝜌𝑒𝑝𝜌Φ@vecv𝑐@vecE@vecB@vecvΠ\frac{\partial(E_{t}+\rho\Phi)}{\partial t}+\nabla\cdot\left[\left(\frac{\rho% \@vec{v}^{2}}{2}+\rho e+p+\rho\Phi\right)\@vec{v}+c\@vec{E}\times\@vec{B}% \right]=\nabla\cdot(\@vec{v}\cdot\Pi),divide start_ARG ∂ ( italic_E start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT + italic_ρ roman_Φ ) end_ARG start_ARG ∂ italic_t end_ARG + ∇ ⋅ [ ( divide start_ARG italic_ρ start_ID start_ARG italic_v end_ARG end_ID start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG + italic_ρ italic_e + italic_p + italic_ρ roman_Φ ) start_ID start_ARG italic_v end_ARG end_ID + italic_c start_ID start_ARG italic_E end_ARG end_ID × start_ID start_ARG italic_B end_ARG end_ID ] = ∇ ⋅ ( start_ID start_ARG italic_v end_ARG end_ID ⋅ roman_Π ) , (4)
(ρs)t+(ρsv)=0𝜌𝑠𝑡𝜌𝑠@vecv0\frac{\partial(\rho s)}{\partial t}+\nabla\cdot\left(\rho s\@vec{v}\right)=0divide start_ARG ∂ ( italic_ρ italic_s ) end_ARG start_ARG ∂ italic_t end_ARG + ∇ ⋅ ( italic_ρ italic_s start_ID start_ARG italic_v end_ARG end_ID ) = 0 (5)

where ρ𝜌\rhoitalic_ρ is the mass density, v@vecv\@vec{v}start_ID start_ARG italic_v end_ARG end_ID is the gas velocity, B@vecB\@vec{B}start_ID start_ARG italic_B end_ARG end_ID is the magnetic field, E@vecE\@vec{E}start_ID start_ARG italic_E end_ARG end_ID is the induced electric field, p𝑝pitalic_p is the gas thermal pressure, s𝑠sitalic_s is the gas entropy, and Et=ρe+m2/2ρ+B2/2subscript𝐸𝑡𝜌𝑒superscript@vecm22𝜌superscript@vecB22E_{t}=\rho e+\@vec{m}^{2}/2\rho+\@vec{B}^{2}/2italic_E start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = italic_ρ italic_e + start_ID start_ARG italic_m end_ARG end_ID start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 italic_ρ + start_ID start_ARG italic_B end_ARG end_ID start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 is the total energy density. The entropy is evolved as a passive scalar since source parabolic terms are always added to the total energy. We treat the fluid as an ideal gas for the closure relation between ρ𝜌\rhoitalic_ρ, e𝑒eitalic_e, and s𝑠sitalic_s, where we take γ=5/3𝛾53\gamma=5/3italic_γ = 5 / 3. The gas pressure is recovered through the entropy, i.e. s=ρ/pγ𝑠𝜌superscript𝑝𝛾s=\rho/p^{\gamma}italic_s = italic_ρ / italic_p start_POSTSUPERSCRIPT italic_γ end_POSTSUPERSCRIPT. ΦΦ\Phiroman_Φ is the external gravitational potential originating from the central protostar, described in Section 2.1.

The induced electric field, E@vecE\@vec{E}start_ID start_ARG italic_E end_ARG end_ID, is

cE=(v+vH)×B+ηcJ,𝑐@vecE@vecvsubscript@vecv𝐻@vecB𝜂𝑐@vecJc\@vec{E}=-(\@vec{v}+\@vec{v}_{H})\times\@vec{B}+\frac{\eta}{c}\cdot\@vec{J},italic_c start_ID start_ARG italic_E end_ARG end_ID = - ( start_ID start_ARG italic_v end_ARG end_ID + start_ID start_ARG italic_v end_ARG end_ID start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT ) × start_ID start_ARG italic_B end_ARG end_ID + divide start_ARG italic_η end_ARG start_ARG italic_c end_ARG ⋅ start_ID start_ARG italic_J end_ARG end_ID , (6)

where vH=J/enesubscript@vecv𝐻@vecJ𝑒subscript𝑛𝑒\@vec{v}_{H}=-\@vec{J}/en_{e}start_ID start_ARG italic_v end_ARG end_ID start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT = - start_ID start_ARG italic_J end_ARG end_ID / italic_e italic_n start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT, e𝑒eitalic_e is the elementary charge, nesubscript𝑛𝑒n_{e}italic_n start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT is the electron number density, J=c×B@vecJ𝑐𝐵\@vec{J}=c\nabla\times Bstart_ID start_ARG italic_J end_ARG end_ID = italic_c ∇ × italic_B is the induced current, and η𝜂\etaitalic_η is the magnetic resistivity. We describe our implementation of the magnetic resistivity, only including Ohmic dissipation, and our prescription for the ionization fraction in Section 2.2. The term vHsubscript@vecv𝐻\@vec{v}_{H}start_ID start_ARG italic_v end_ARG end_ID start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT introduces the Hall effect into the non-ideal MHD equations. We include shear viscosity, νssubscript𝜈𝑠\nu_{s}italic_ν start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, which is coupled to the above equations with the viscous stress tensor

Π=ηs[v+(v)T]23ηs(v)𝕀.Πsubscript𝜂𝑠delimited-[]@vecvsuperscript@vecv𝑇23subscript𝜂𝑠@vecv𝕀\Pi=\eta_{s}\left[\nabla\@vec{v}+(\nabla\@vec{v})^{T}\right]-\frac{2}{3}\eta_{% s}\left(\nabla\cdot\@vec{v}\right)\mathbb{I}.roman_Π = italic_η start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT [ ∇ start_ID start_ARG italic_v end_ARG end_ID + ( ∇ start_ID start_ARG italic_v end_ARG end_ID ) start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT ] - divide start_ARG 2 end_ARG start_ARG 3 end_ARG italic_η start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( ∇ ⋅ start_ID start_ARG italic_v end_ARG end_ID ) blackboard_I . (7)

We solve these equations in axisymmetric spherical coordinates (r, θ𝜃\thetaitalic_θ), including all three components of vectors (2.5D). We use 512 radial logarithmically spaced cells from the protostellar surface out to 1 AU. Along the θ𝜃\thetaitalic_θ direction, we use 512 cells between [102,π102]superscript102𝜋superscript102[10^{-2},\pi-10^{-2}][ 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT , italic_π - 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT ] radians to avoid the poles at θ=0,π𝜃0𝜋\theta=0,\piitalic_θ = 0 , italic_π. This leads to Δr1×104Δ𝑟1superscript104\Delta r\approx 1\times 10^{-4}roman_Δ italic_r ≈ 1 × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT AU at the protostar surface and Δr8×103Δ𝑟8superscript103\Delta r\approx 8\times 10^{-3}roman_Δ italic_r ≈ 8 × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT AU at the 1 AU boundary. The high resolution near the protostellar surface is essential to capture the accretion flow structure and turbulence, which is lost at lower resolution. The inner bloated protostar surface is R=3R=0.01395subscript𝑅3subscriptRdirect-product0.01395R_{*}=3{\,\rm R_{\odot}}=0.01395italic_R start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT = 3 roman_R start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT = 0.01395 AU, motivated by models accreting bloating solar mass protostars (Geroux et al., 2016)),

2.1 Initial Conditions

We follow the convention that (r,θ,ϕ)𝑟𝜃italic-ϕ(r,\leavevmode\nobreak\ \theta,\leavevmode\nobreak\ \phi)( italic_r , italic_θ , italic_ϕ ) denotes the spherical coordinates, and the two-dimensional cylindrical coordinates are (R=rsinθ,z=rcosθ)formulae-sequence𝑅𝑟𝜃𝑧𝑟𝜃(R=r\sin\theta,\leavevmode\nobreak\ z=r\cos\theta)( italic_R = italic_r roman_sin italic_θ , italic_z = italic_r roman_cos italic_θ ). The initial disk surface density profile is the often-used truncated-power law (e.g. Lynden-Bell & Pringle, 1974; Béthune et al., 2017; Zhu et al., 2024; Rea et al., 2024). We assume an initial disk surface density given by a purely viscous disk,

Σ(R)=Σ0(RR0)1eR/Rmax,Σ𝑅subscriptΣ0superscript𝑅subscript𝑅01superscript𝑒𝑅subscript𝑅max\Sigma(R)=\Sigma_{0}\left(\frac{R}{R_{0}}\right)^{-1}e^{-R/R_{\rm max}},roman_Σ ( italic_R ) = roman_Σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( divide start_ARG italic_R end_ARG start_ARG italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - italic_R / italic_R start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT end_POSTSUPERSCRIPT , (8)

where ΣΣ\Sigmaroman_Σ is the gas surface density, Σ0subscriptΣ0\Sigma_{0}roman_Σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the normalizing factor, described below, R𝑅Ritalic_R is the cylindrical radius and R0subscript𝑅0R_{0}italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the normalization radius, taken to be R0=10subscript𝑅010R_{0}=10italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 10 AU. The initial disk is assumed to be in hydrostatic balance, such that the gas density is

ρ=Σ(R)2πhexp(z22h2)𝜌Σ𝑅2𝜋superscript𝑧22superscript2\rho=\frac{\Sigma(R)}{\sqrt{2\pi}h}\exp{\left(-\frac{z^{2}}{2h^{2}}\right)}italic_ρ = divide start_ARG roman_Σ ( italic_R ) end_ARG start_ARG square-root start_ARG 2 italic_π end_ARG italic_h end_ARG roman_exp ( - divide start_ARG italic_z start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_h start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) (9)

where h=cs/ΩKsubscript𝑐𝑠subscriptΩ𝐾h=c_{s}/\Omega_{K}italic_h = italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT / roman_Ω start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT is the scale height of the disk and ΩK=Gm/R3subscriptΩ𝐾𝐺subscript𝑚superscript𝑅3\Omega_{K}=\sqrt{Gm_{*}/R^{3}}roman_Ω start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT = square-root start_ARG italic_G italic_m start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT / italic_R start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG is the Keplerian rotation rate. The disk extends out to Rmax=40subscript𝑅max40R_{\rm max}=40italic_R start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 40 AU, consistent with survey results of protostars in the Orion molecular cloud similar to our model (Tobin et al., 2020; Sheehan et al., 2022). Our initial disk surface density profile is marginally steeper than the average of the modeled disks in Sheehan et al. (2022) but is still within the range of their survey. The tangential velocity is the Keplerian speed

vϕ=GmR.subscript𝑣italic-ϕ𝐺subscript𝑚𝑅v_{\phi}=\sqrt{\frac{Gm_{*}}{R}}.italic_v start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT = square-root start_ARG divide start_ARG italic_G italic_m start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT end_ARG start_ARG italic_R end_ARG end_ARG . (10)

There is a prescribed initial radial velocity as a function of the target accretion rate, m˙subscript˙𝑚\dot{m}_{*}over˙ start_ARG italic_m end_ARG start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT, assuming the accretion onto the surface is the same as through the disk, i.e. m˙=m˙disksubscript˙𝑚subscript˙𝑚disk\dot{m}_{*}=\dot{m}_{\rm disk}over˙ start_ARG italic_m end_ARG start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT = over˙ start_ARG italic_m end_ARG start_POSTSUBSCRIPT roman_disk end_POSTSUBSCRIPT. The radial velocity is set by mass conservation through the disk, i.e.,

vr=m˙disk2πΣ(R)R.subscript𝑣𝑟subscript˙𝑚disk2𝜋Σ𝑅𝑅v_{r}=-\frac{\dot{m}_{\rm disk}}{2\pi\Sigma(R)R}.italic_v start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT = - divide start_ARG over˙ start_ARG italic_m end_ARG start_POSTSUBSCRIPT roman_disk end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_π roman_Σ ( italic_R ) italic_R end_ARG . (11)

Finally, there is an initial core described by a purely infalling envelope (Ulrich, 1976), with a density profile

ρenv=ρ0,env(r50AU)1.5.subscript𝜌envsubscript𝜌0envsuperscript𝑟50AU1.5\rho_{\rm env}=\rho_{\rm 0,env}\left(\frac{r}{50\,{\rm AU}}\right)^{-1.5}.italic_ρ start_POSTSUBSCRIPT roman_env end_POSTSUBSCRIPT = italic_ρ start_POSTSUBSCRIPT 0 , roman_env end_POSTSUBSCRIPT ( divide start_ARG italic_r end_ARG start_ARG 50 roman_AU end_ARG ) start_POSTSUPERSCRIPT - 1.5 end_POSTSUPERSCRIPT . (12)

While the initial core is disrupted in the simulation by outflowing gas, it can remain in the outer domain and helps balance the pressure of the disk in the initial condition.

The values of Σ0subscriptΣ0\Sigma_{0}roman_Σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and ρ0,envsubscript𝜌0env\rho_{\rm 0,env}italic_ρ start_POSTSUBSCRIPT 0 , roman_env end_POSTSUBSCRIPT are determined to get a total mass ratio between the protostar and its natal disk and core, such that mdisk/m=0.175subscript𝑚disksubscript𝑚0.175m_{\rm disk}/m_{*}=0.175italic_m start_POSTSUBSCRIPT roman_disk end_POSTSUBSCRIPT / italic_m start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT = 0.175 for the whole disk (out to Rmax=40subscript𝑅max40R_{\rm max}=40italic_R start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 40 AU) and mcore/m=0.1subscript𝑚coresubscript𝑚0.1m_{\rm core}/m_{*}=0.1italic_m start_POSTSUBSCRIPT roman_core end_POSTSUBSCRIPT / italic_m start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT = 0.1 (out to 0.1 pc).

The initial temperature is assumed to vary with cylindrical radius as

T(r)=T0(RR0)0.5,𝑇𝑟subscript𝑇0superscript𝑅subscript𝑅00.5T(r)=T_{0}\left(\frac{R}{R_{0}}\right)^{-0.5},italic_T ( italic_r ) = italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( divide start_ARG italic_R end_ARG start_ARG italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT - 0.5 end_POSTSUPERSCRIPT , (13)

where T0=200subscript𝑇0200T_{0}=200\>italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 200K. This temperature is warmer than the recommended relation from Sheehan et al. (2022) using only the intrinsic protostellar luminosity, but it is consistent if the accretion luminosity corresponding to the targetted accretion rate is added as an additional luminosity component. The temperature radial scaling exponent can take a range of values (Kenyon et al., 1993; Chiang & Goldreich, 1997), and a value of 0.50.5-0.5- 0.5 is consistent with radiative transfer modeling of embedded protostellar disks Tobin et al. (2020).

Both the protostar and the disk are magnetized. We define an initial disk magnetic field similar to Zanni & Ferreira (2009), such that the magnetic field is vertical in the outer disk and transitions to an hourglass morphology in the inner disk. The magnetic field is initialized by the vector potential, A@vecA\@vec{A}start_ID start_ARG italic_A end_ARG end_ID, where B=×A@vecB@vecA\@vec{B}=\nabla\times\@vec{A}start_ID start_ARG italic_B end_ARG end_ID = ∇ × start_ID start_ARG italic_A end_ARG end_ID. Given the axisymmetry, we only need to initialize the ϕitalic-ϕ\phiitalic_ϕ-component:

Aϕ=2B1R13ar(a1)/2,subscript𝐴italic-ϕ2subscript𝐵1subscript𝑅13𝑎superscript𝑟𝑎12A_{\phi}=\frac{2B_{1}R_{1}}{3-a}r^{-(a-1)/2},italic_A start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT = divide start_ARG 2 italic_B start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG 3 - italic_a end_ARG italic_r start_POSTSUPERSCRIPT - ( italic_a - 1 ) / 2 end_POSTSUPERSCRIPT , (14)

where B1=2p1/β1subscript𝐵12subscript𝑝1subscript𝛽1B_{1}=\sqrt{2p_{1}/\beta_{1}}italic_B start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = square-root start_ARG 2 italic_p start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT / italic_β start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG, p1subscript𝑝1p_{1}italic_p start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT is the pressure at 1 AU, R1=1subscript𝑅11R_{1}=1italic_R start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 1 AU, and β1=p1/p1,magsubscript𝛽1subscript𝑝1subscript𝑝1mag\beta_{1}=p_{1}/p_{\rm 1,mag}italic_β start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_p start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT / italic_p start_POSTSUBSCRIPT 1 , roman_mag end_POSTSUBSCRIPT is plasma beta, the ratio of thermal pressure to magnetic pressure, calculated at 1 AU. We use a=1𝑎1a=1italic_a = 1 as the simplest initial case. This produces an initial poloidal magnetic field which has an hourglass morphology in the inner regions and transitions to a vertical field by R0.5𝑅0.5R\approx 0.5italic_R ≈ 0.5 AU. We use β=105𝛽superscript105\beta=10^{5}italic_β = 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT, which leads to a magnetic field at the pole of \approx 200 G. The protostellar magnetic field is included as a background dipole magnetic field, derived by the magnetic vector potential

Aϕ,=BR3r2,subscript𝐴italic-ϕsubscript𝐵superscriptsubscript𝑅3superscript𝑟2A_{\phi,*}=\frac{B_{*}R_{*}^{3}}{r^{2}},italic_A start_POSTSUBSCRIPT italic_ϕ , ∗ end_POSTSUBSCRIPT = divide start_ARG italic_B start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG start_ARG italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG , (15)

where Bsubscript𝐵B_{*}italic_B start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT is the surface magnetic field strength. We explore four different protostellar magnetic fields, with B=subscript𝐵absentB_{*}=italic_B start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT =10, 500, 1000, and 2000 G fields, encompassing the range of observationally inferred surface magnetic fields in the recent survey of low-mass protostars from Flores et al. (2024). Evolved protostars, in particular T-Tauri stars, have been found to have complex magnetic fields (Donati et al., 2007, 2008; Gregory et al., 2008; Carroll et al., 2012), but there are no constraints on this for embedded protostars. Therefore, we assume a pure dipole, typical of many studies (e.g. Rosen et al., 2012; Zanni & Ferreira, 2013; Ireland et al., 2021; Takasao et al., 2022), but note that the disk magnetic field will perturb the field around the star, regardless, away from a pure dipole.

We include the protostar’s gravitational potential following

Φ(r)=Gmr,Φ𝑟𝐺subscript𝑚𝑟\Phi(r)=-\frac{Gm_{*}}{r},roman_Φ ( italic_r ) = - divide start_ARG italic_G italic_m start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT end_ARG start_ARG italic_r end_ARG , (16)

such that the gravitational acceleration due to the protostar is agrav=Φsubscript@vecagravΦ\@vec{a}_{\rm grav}=-\nabla\Phistart_ID start_ARG italic_a end_ARG end_ID start_POSTSUBSCRIPT roman_grav end_POSTSUBSCRIPT = - ∇ roman_Φ. We neglect the disk self-gravity since in this inner 1 AU, over these short timescales, the gravitational potential will be dominated by the central protostar.

The protostar, which is inside the inner boundary, is assumed to be rotating at some fraction, fbusubscript𝑓buf_{\rm bu}italic_f start_POSTSUBSCRIPT roman_bu end_POSTSUBSCRIPT, of the breakup speed,

Ω=fbuGmR3,subscriptΩsubscript𝑓bu𝐺subscript𝑚superscriptsubscript𝑅3\Omega_{*}=f_{\rm bu}\sqrt{\frac{Gm_{*}}{R_{*}^{3}}},roman_Ω start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT = italic_f start_POSTSUBSCRIPT roman_bu end_POSTSUBSCRIPT square-root start_ARG divide start_ARG italic_G italic_m start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT end_ARG start_ARG italic_R start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG end_ARG , (17)

where we take fbu=1/3subscript𝑓bu13f_{\rm bu}=1/3italic_f start_POSTSUBSCRIPT roman_bu end_POSTSUBSCRIPT = 1 / 3 as a nomial median value between slow- and fast-rotators (White & Hillenbrand, 2004; Herbst et al., 2007; Rosen et al., 2012). The value of fbu=1/3subscript𝑓bu13f_{\rm bu}=1/3italic_f start_POSTSUBSCRIPT roman_bu end_POSTSUBSCRIPT = 1 / 3 gives a rotation rate of 4.026×1054.026superscript1054.026\times 10^{-5}4.026 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT s-1, or a rotation period of Trot=2π/Ω1.56×105subscript𝑇rot2𝜋subscriptΩ1.56superscript105T_{\rm rot}=2\pi/\Omega_{*}\approx 1.56\times 10^{5}italic_T start_POSTSUBSCRIPT roman_rot end_POSTSUBSCRIPT = 2 italic_π / roman_Ω start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT ≈ 1.56 × 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT seconds. The actual rotation speeds of protostars are largely unknown and will be explored in future simulation suites, in particular for highly magnetized protostars.

2.2 Viscosity and Resistivity

The shear viscosity is included through an α𝛼\alphaitalic_α-viscosity parameterization (Shakura & Sunyaev, 1973),

ηs=αvcsh,subscript𝜂𝑠subscript𝛼𝑣subscript𝑐𝑠\eta_{s}=\alpha_{v}c_{s}h,italic_η start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = italic_α start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT italic_h , (18)

where cssubscript𝑐𝑠c_{s}italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT is the sound speed and hhitalic_h is the scale height, computed by using the local sound speed and Keplerian rotation rate. We set αv=0.01subscript𝛼𝑣0.01\alpha_{v}=0.01italic_α start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT = 0.01, which for our initial setup would correspond to a floor accretion rate of m˙107˙𝑚superscript107\dot{m}\approx 10^{-7}over˙ start_ARG italic_m end_ARG ≈ 10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT M yr-1. This effectively acts as a viscosity minimum, to ensure a minimal degree of accretion and angular momentum loss through the disk. However, at our simulation resolutions, it is possible to resolve turbulence within the inner disk which will act as an additional source of angular momentum loss.

For the non-ideal magnetic resistivity, we include Ohmic dissipation following Balbus & Terquem (2001); Lesur et al. (2014),

ηO=250xeTcms2,subscript𝜂O250subscript𝑥𝑒𝑇cmsuperscripts2\eta_{\rm O}=\frac{250}{x_{e}}\sqrt{T}{\>\rm cm\,s^{-2}},italic_η start_POSTSUBSCRIPT roman_O end_POSTSUBSCRIPT = divide start_ARG 250 end_ARG start_ARG italic_x start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT end_ARG square-root start_ARG italic_T end_ARG roman_cm roman_s start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT , (19)

where xesubscript𝑥𝑒x_{e}italic_x start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT is the ionization fraction. We compute the ionization fraction assuming two components: thermal ionization, which will be dominated by low ionization potential alkalis Na+ and K+ (Balbus & Hawley, 2000), and X-ray ionization. The electron fraction due to the thermal ionization of Potassium is (Balbus & Hawley, 2000; Fromang et al., 2002),

xeth6.47×1013(xK107)(T103)0.75×(2.4×1015cm3n)(e(25188K)/T1.15×1011),superscriptsubscript𝑥𝑒th6.47superscript1013subscript𝑥𝐾superscript107superscript𝑇superscript1030.752.4superscript1015superscriptcm3𝑛superscript𝑒25188K𝑇1.15superscript1011x_{e}^{\rm th}\approx 6.47\times 10^{-13}\left(\frac{x_{K}}{10^{-7}}\right)% \left(\frac{T}{10^{3}}\right)^{0.75}\times\\ \left(\frac{2.4\times 10^{15}{\>\rm cm^{-3}}}{n}\right)\left(\frac{e^{-(25188{% \>\rm K})/T}}{1.15\times 10^{-11}}\right),start_ROW start_CELL italic_x start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_th end_POSTSUPERSCRIPT ≈ 6.47 × 10 start_POSTSUPERSCRIPT - 13 end_POSTSUPERSCRIPT ( divide start_ARG italic_x start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT end_ARG start_ARG 10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT end_ARG ) ( divide start_ARG italic_T end_ARG start_ARG 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG ) start_POSTSUPERSCRIPT 0.75 end_POSTSUPERSCRIPT × end_CELL end_ROW start_ROW start_CELL ( divide start_ARG 2.4 × 10 start_POSTSUPERSCRIPT 15 end_POSTSUPERSCRIPT roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT end_ARG start_ARG italic_n end_ARG ) ( divide start_ARG italic_e start_POSTSUPERSCRIPT - ( 25188 roman_K ) / italic_T end_POSTSUPERSCRIPT end_ARG start_ARG 1.15 × 10 start_POSTSUPERSCRIPT - 11 end_POSTSUPERSCRIPT end_ARG ) , end_CELL end_ROW (20)

where xKsubscript𝑥𝐾x_{K}italic_x start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT is the potassium abundance, taken to be xK=107subscript𝑥𝐾superscript107x_{K}=10^{-7}italic_x start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT (Fromang et al., 2002), and n𝑛nitalic_n is the gas hydrogen-nuclei number density.

For the X-ray ionization, we break the domain into two components using the initial disk density profile. For regions outside of the disk, defined by z/h>4𝑧4z/h>4italic_z / italic_h > 4, we assume the X-ray flux is optically thin,

FX(r)=LX4πr2,subscript𝐹𝑋𝑟subscript𝐿𝑋4𝜋superscript𝑟2F_{X}(r)=\frac{L_{X}}{4\pi r^{2}},italic_F start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT ( italic_r ) = divide start_ARG italic_L start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT end_ARG start_ARG 4 italic_π italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG , (21)

where LXsubscript𝐿𝑋L_{X}italic_L start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT is an assumed X-ray luminosity, nominally taken to be LX=1030subscript𝐿𝑋superscript1030L_{X}=10^{30}italic_L start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 30 end_POSTSUPERSCRIPT erg s-1, consistent with a solar mass protostar (Getman et al., 2022). Inside the disk, we attenuate the X-ray emission vertically through the disk using a column density computed by the initial disk profile,

FX(r)=FX(R,z)=LX4πR2eσX0zn(R,z)𝑑z,subscript𝐹𝑋𝑟subscript𝐹𝑋𝑅𝑧subscript𝐿𝑋4𝜋superscript𝑅2superscript𝑒subscript𝜎𝑋superscriptsubscript0𝑧𝑛𝑅superscript𝑧differential-dsuperscript𝑧F_{X}(r)=F_{X}(R,z)=\frac{L_{X}}{4\pi R^{2}}e^{-\sigma_{X}\int_{0}^{z}n(R,z^{% \prime})dz^{\prime}},italic_F start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT ( italic_r ) = italic_F start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT ( italic_R , italic_z ) = divide start_ARG italic_L start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT end_ARG start_ARG 4 italic_π italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG italic_e start_POSTSUPERSCRIPT - italic_σ start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT italic_n ( italic_R , italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) italic_d italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT , (22)

where n(R,z)𝑛𝑅𝑧n(R,z)italic_n ( italic_R , italic_z ) is the number density of the initial disk density distribution, and σX1022subscript𝜎𝑋superscript1022\sigma_{X}\approx 10^{-22}italic_σ start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT ≈ 10 start_POSTSUPERSCRIPT - 22 end_POSTSUPERSCRIPT cm-2 is the X-ray photo-absorption cross section around 1 keV for gas with dominantly neutral hydrogen (Gaches et al., 2023) (does not vary much until the gas temperature greatly exceeds 104superscript10410^{4}10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT K). Since the initial disk is assumed to be in hydrostatic equilibrium, the integral in the exponential has an analytic solution. The X-ray ionization rate, ζ𝜁\zetaitalic_ζ, is computed assuming that it is dominated by secondary ionization, determined by the energy deposition (see below) and the mean energy per ion pair (Dalgarno et al., 1999; Meijerink & Spaans, 2005).

The ionization fraction due to X-ray irradiation is then

xeζ=ζβn,superscriptsubscript𝑥𝑒𝜁𝜁𝛽𝑛x_{e}^{\zeta}=\sqrt{\frac{\zeta}{\beta n}},italic_x start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_ζ end_POSTSUPERSCRIPT = square-root start_ARG divide start_ARG italic_ζ end_ARG start_ARG italic_β italic_n end_ARG end_ARG , (23)

where β𝛽\betaitalic_β is the recombination rate for molecular ions (Fromang et al., 2002). By construction, ζ𝜁\zetaitalic_ζ is constant throughout the disk, but xeζsuperscriptsubscript𝑥𝑒𝜁x_{e}^{\zeta}italic_x start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_ζ end_POSTSUPERSCRIPT will vary due to the changing gas density. The total ionization fraction is xe=xeth+xeζsubscript𝑥𝑒superscriptsubscript𝑥𝑒thsuperscriptsubscript𝑥𝑒𝜁x_{e}=x_{e}^{\rm th}+x_{e}^{\zeta}italic_x start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT = italic_x start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_th end_POSTSUPERSCRIPT + italic_x start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_ζ end_POSTSUPERSCRIPT. The electron density is ne=xensubscript𝑛𝑒subscript𝑥𝑒𝑛n_{e}=x_{e}nitalic_n start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT = italic_x start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT italic_n. Including a full self-consistent ionization solver is not within the scope of this work, which focuses on the impact of the protostellar magnetic field on the accretion flow.

We only include the viscosity and resistivity where rrcorot𝑟subscript𝑟corotr\geq r_{\rm corot}italic_r ≥ italic_r start_POSTSUBSCRIPT roman_corot end_POSTSUBSCRIPT, where rcorot=(GMΩ2)1/3subscript𝑟corotsuperscript𝐺subscript𝑀superscriptsubscriptΩ213r_{\rm corot}=\left(\frac{GM_{*}}{\Omega_{*}^{2}}\right)^{1/3}italic_r start_POSTSUBSCRIPT roman_corot end_POSTSUBSCRIPT = ( divide start_ARG italic_G italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT end_ARG start_ARG roman_Ω start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) start_POSTSUPERSCRIPT 1 / 3 end_POSTSUPERSCRIPT is the corotation radius since inside this radius the gas is expected to be ionized enough for the gas to evolve more similarly to ideal MHD. We taper the resistivity and viscosity to zero from the inner disk boundary to rcorot=0.029subscript𝑟corot0.029r_{\rm corot}=0.029italic_r start_POSTSUBSCRIPT roman_corot end_POSTSUBSCRIPT = 0.029 AU (for our assumed fbusubscript𝑓buf_{\rm bu}italic_f start_POSTSUBSCRIPT roman_bu end_POSTSUBSCRIPT) to avoid a discontinuity in the viscosity and resistivity.

2.3 Heating and cooling

We include both atomic line and continuum cooling for gas between 104K<T<109Ksuperscript104K𝑇superscript109K10^{4}{\,\rm K}<T<10^{9}{\,\rm K}10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT roman_K < italic_T < 10 start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT roman_K and X-ray gas heating from the central protostar. The line cooling is included as a tabulated cooling rate calculated using the ChiantiPy (Dere, 2013; Dere et al., 2019) package for solar abundances and collisional ionization equilibrium. Since we do not include chemistry, we neglect molecular line cooling. Further, we do not include dust cooling, and as such the simulations may be warmer, especially in the disk where the gas temperature can drop below the dust sublimation temperature. X-ray gas heating is included using a simple prescription. The heating term is

εX=nHHXergs1cm3,subscript𝜀𝑋subscript𝑛𝐻subscript𝐻𝑋ergsuperscripts1superscriptcm3\varepsilon_{X}=n_{H}H_{X}\,\,{\rm erg\,s^{-1}\,cm^{-3}},italic_ε start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT = italic_n start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT italic_H start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT roman_erg roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT , (24)

where HXFXσXsubscript𝐻𝑋subscript𝐹𝑋subscript𝜎𝑋H_{X}\approx F_{X}\sigma_{X}italic_H start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT ≈ italic_F start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT is the X-ray energy deposition rate and FXsubscript𝐹𝑋F_{X}italic_F start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT is the X-ray flux computed as described above. X-ray heating is necessary to ensure a heated envelope above the protostar and a smoother thermal profile away from the protostar. In the non-ideal, viscous MHD equations, there are also heating contributions due to the dissipation of the magnetic field and shear viscosity.

2.4 Boundary conditions

The inner boundary must be treated carefully to avoid inducing artifacts in the magnetic field from the rotating, magnetized protostar surface (Zanni & Ferreira, 2009). We assume that the protostar is a perfect conductor, such that E=B×(uΩ×R)=0@vecE@vecB@vecu@vecΩ@vecR0\@vec{E}=\@vec{B}\times(\@vec{u}-\@vec{\Omega_{*}}\times\@vec{R})=0start_ID start_ARG italic_E end_ARG end_ID = start_ID start_ARG italic_B end_ARG end_ID × ( start_ID start_ARG italic_u end_ARG end_ID - start_ID start_ARG roman_Ω start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT end_ARG end_ID × start_ID start_ARG italic_R end_ARG end_ID ) = 0. The protostar is set to be rotating aligned with the θ=0𝜃0\theta=0italic_θ = 0 pole. For infalling material, we force the condition that Eϕ=0subscript𝐸italic-ϕ0E_{\phi}=0italic_E start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT = 0 by setting the poloidal velocity components according to the poloidal magnetic field. For outflowing material, we set the invariant k=4πρup/Bp𝑘4𝜋𝜌subscript𝑢𝑝subscript𝐵𝑝k=4\pi\rho u_{p}/B_{p}italic_k = 4 italic_π italic_ρ italic_u start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT / italic_B start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT to be constant along the radial boundary. The Brsubscript𝐵𝑟B_{r}italic_B start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT and Bθsubscript𝐵𝜃B_{\theta}italic_B start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT components use an outflow boundary condition. To ensure we do not induce an electron field inside the magnetic field, we set the toroidal magnetic field by assuming RBϕ𝑅subscript𝐵italic-ϕRB_{\phi}italic_R italic_B start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT is conserved across the boundary and that uϕ=RΩ+upBϕ/Bpsubscript𝑢italic-ϕ𝑅subscriptΩsubscript𝑢𝑝subscript𝐵italic-ϕsubscript𝐵𝑝u_{\phi}=R\Omega_{*}+u_{p}B_{\phi}/B_{p}italic_u start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT = italic_R roman_Ω start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT + italic_u start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT italic_B start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT / italic_B start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT. These criteria ensure that the gas falling onto the protostellar sink surface is absorbed perfectly through the boundary. At the outer radial boundary, we assume an outflow-but-no-inflow condition. We use axisymmetric boundary conditions in the polar direction.

3 Results

The high spatial resolution, from 104102absentsuperscript104superscript102\approx 10^{-4}-10^{-2}≈ 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT AU, and high-temporal output cadence, \approx30 minutes, provides a unique view into the role the protostellar magnetic field plays in the underlying accretion physics. The simulations are evolved for 60 days due to timestep constraints caused by the highest wave speeds tending to be in the regions with the highest resolution and to avoid any noticeable disk depletion caused by the lack of inflow at the outer boundary. For the analysis presented below, we ignore the first 11 days of evolution to remove the impact of transients from the relaxation of the initial conditions. This results in 49 days, or approximately 27 protostellar rotation periods, defined by Trot=2π/Ωsubscript𝑇rot2𝜋subscriptΩT_{\rm rot}=2\pi/\Omega_{*}italic_T start_POSTSUBSCRIPT roman_rot end_POSTSUBSCRIPT = 2 italic_π / roman_Ω start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT, of evolution after the transient. In the following, we discuss the morphology of the accretion disk, the differences in the accretion variability depending on the strength of the protostellar magnetic field, and examine the accretion flow onto the star and determine its impact location and covering fraction on the protostellar surface.

3.1 The inner 1 AU: impact of protostellar magnetic field

Figure 1 shows the density distribution of the gas from 1 AU scales down to the protostellar surface, with the insets showing progressively higher zoom-ins. The density profiles of the four disks are morphologically similar at the outer radii, as shown in Figure 2, which displays the density distribution averaged 22 around the midplane. This suggests that the disk’s bulk is likely influenced more by the disk magnetic field and viscosity effects than by the protostellar magnetic fields. Overall, the disks appear to be lifted, which is expected due to the magnetic buoyancy pulling gas from the disk out as a wind. The upper disk layers and outflow cavities however show significant differences between each other. While all simulations feature outflows, the nature of these appears very different. For the low magnetized protostar, the outflow appears very turbulent, dominated by gas which is thrown off the protostar near the surface. The more magnetized protostars exhibit outflows that are more collimated and ordered, with filaments/knots of gas from magnetospheric ejections near the protostellar surface. Moreover, we also find one-sided outflows, which are caused by breaking the symmetry of the protostellar dipole magnetic field due to its interaction with the disk magnetic field. Such one-sided outflows have been observed in some nearby protostars (e.g., Federman et al., 2024) and simulations of T-Tauri-like protostars with complex fields (Lovelace et al., 2010).

Refer to caption Refer to caption
Refer to caption Refer to caption
Figure 1: Density distribution for the 10 Gauss (top left), 500 Gauss (top right), 1000 Gauss (bottom left), and 2000 Gauss (bottom right) protostar, with zoom in varying the box size from 2 AU to 0.04 AU.
Refer to caption
Figure 2: Gas density, ρmidsubscript𝜌mid\rho_{\rm mid}italic_ρ start_POSTSUBSCRIPT roman_mid end_POSTSUBSCRIPT versus cylindrical radius, R𝑅Ritalic_R, averaged 22 around the midplane, at the final simulation snapshot for the four different protostar magnetic field models, demonstrating that the out disk midplanes are not impacted by the protostellar magnetic field.

As we zoom in, the figure shows substantial differences begin to appear between the different protostar magnetic field simulations. The 10 G protostar exhibits a highly turbulent flow within 0.1 AU, with the inflow covering the vast fraction of the protostellar surface. The very weak protostellar dipole means there is substantially less magnetic pressure to counter the ram pressure of the infalling gas. The 500 G, 1 kG, and 2 kG protostars exhibit a more ordered inner 0.1 AU, with a clear differentiation between the magnetically driven bipolar outflow cavity and the accreting gas. For the 1 kG and 2 kG protostar fields, the disk is truncated due to the protostar’s magnetic pressure causing gas to flow along the protostar’s magnetic field lines where it is funneled to the star in agreement with previous work (Koenigl, 1991). These most inner regions can traced by hydrogen recombination emission, particularly in the NIR, hinting that probing these regions with high-resolution spectroscopy could provide a direct test of the accretion physics, as is done for more evolved objects (e.g. Gravity Collaboration et al., 2023b, a).

3.2 Where does accretion occur?

Where accretion occurs is of substantial interest in star formation physics. Star formation simulations utilize sink particles with sub-grid protostellar evolution models (e.g. Offner et al., 2009) that predict the quantity of gas accreted, but not how and where. Sub-grid models often utilize ad-hoc prescriptions for the accretion luminosity and launched outflows (e.g. Cunningham et al., 2011, 2018). The physical mechanism, covering fraction, and location of the accretion flow onto the surface are also important assumptions utilized in radiation transfer models of protostar accretion (see Hartmann et al., 2016, and references therein).

Under the canonical boundary layer and magnetospheric accretion models, one would find that the accretion flow is either dominantly at the equator, or in high latitude accretion columns, respectively. Figure 3 summarizes where accretion occurs onto the protostar, via the accretion momentum density (or accretion rate per area, ρvr𝜌subscript𝑣𝑟\rho v_{r}italic_ρ italic_v start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT). As the protostellar magnetic field increases, the accretion behavior changes in a non-monotonic way. For low protostellar magnetic fields (10 G), the gas falls across the entire surface in a highly chaotic manner, dominated by a filament impacting the equator. At moderate protostellar magnetic fields (500 G), the disk still impacts the protostellar surface directly, with the bipolar outflow providing pressure inhibiting gas flow to high latitudes. Finally, at high protostellar magnetic fields (1 kG and 2 kG), the flow becomes funneled through accretion columns along magnetic field lines, with one accretion column dominating.

Figure 4 shows the evolution of time-averaged angular probability distributions of accretion onto the protostar for the different protostellar magnetic fields, shown in different 5-day blocks. The probability is averaging the angular distribution for each snapshot, i𝑖iitalic_i, in the block, Pi(θ)m˙(θ)/m˙subscript𝑃𝑖𝜃˙𝑚𝜃˙𝑚P_{i}(\theta)\approx\dot{m}(\theta)/\dot{m}italic_P start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_θ ) ≈ over˙ start_ARG italic_m end_ARG ( italic_θ ) / over˙ start_ARG italic_m end_ARG, where m˙(θ)˙𝑚𝜃\dot{m}(\theta)over˙ start_ARG italic_m end_ARG ( italic_θ ) is the accretion rate across the surface. The block’s central time is annotated. Since the distributions are all normalized to the total gas accretion in each block, we can directly compare the differences between the different magnetic fields and their accretion distribution evolution.

Even from early times, there are substantial systematic differences between the angular accretion distributions between the different magnetic fields. The 10 G protostar field shows a distribution that varies substantially between the time blocks, with periods where the bulk of the gas is accreted at the equator and other blocks where the gas is accreted more evenly distributed across the surface. The higher latitude accretion peaks in the 10 G model are from failed winds falling back along the disk surface. For the 500 G protostar field, the accretion occurs almost entirely as an equatorial flow. The accretion flow is constrained by the strong bipolar outflow, resulting in almost all the accretion occurring within ±30plus-or-minussuperscript30\pm 30^{\circ}± 30 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT. At higher protostellar magnetic fields (1 kG and 2 kG) the disk becomes truncated, with an accretion column impacting the protostar broadly at higher latitudes, although in different hemispheres. Lower latitude flows and equatorial accretion occur in the 1 kG and 2 kG simulations during accretion bursts from matter streaming directly from the truncation radius to the protostar (see Section 3.4).

Refer to caption
Figure 3: Accretion rates per unit area across the protostellar surface as a function of time for the four difference cases.
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 4: Probability distributions of the angles (in degrees) at which mass is accreted onto the central protostar over \approx 5-day blocks for the 10 G (left) 500 G (middle left) and 1000 Gauss (middle right) and 2000 Gauss (right) protostars. Annotated is the block number (also denoted by color), and the time in the center of the block in days.

3.3 Accretion Variability and Morphology Dependence on Bsubscript𝐵B_{*}italic_B start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT

The evolution of the accretion rate over time, including any periodic variability and stochasticity, is a potential diagnostic of accretion physics. Figure 5 shows the total accretion rate, m˙˙𝑚\dot{m}over˙ start_ARG italic_m end_ARG, in Msubscript𝑀direct-productM_{\odot}italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT yr-1, for the different simulations. Figure 5 shows both the instantaneous accretion rate and a day-long rolling average. The simulations exhibit rapid accretion rate spikes sometimes over an order of magnitude due to the turbulent nature of the inner accretion flow. Since the gas flows in the inner disk are stochastic, the gas hitting the surface can occur in clumps, boosting the accretion rate up substantially between each output. However, as shown by the black dashed lines in the figure, the 1-day rolling averages show substantially less variability for the 10 G, 500 G, and 1 kG models. There is substantial systematic variability in the 2 kG model since its accretion is dominated by a bursting mode.

Refer to caption
Figure 5: Top four panels: Accretion rate in Msubscript𝑀direct-productM_{\odot}italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT yr-1 versus time for the four different protostar magnetic fields, with the rolling average over a day-long window in cyan. Bottom: 1-day rolling average accretion rates for the different protostellar magnetic fields.

There are clear differences in the accretion rate evolution between the four different cases. Most importantly, the variability of the accretion flow onto the protostar is dependent on the protostar’s magnetic field strength. In the lowly magnetized protostar, accretion occurs often via clumps in the turbulent flow. As the protostar’s magnetic field is increased, the magnitude of the spikes decreases, as does their occurrence rate. This is due to the protostellar magnetic field launching an outflow and the strong ordered field reducing the turbulence in the accretion region. Therefore, the rapid variability decreases with increasing magnetic field strength. However, for the strongest magnetic field models, 1 kG and 2 kG, a more systematic variability begins occurring. At 2 kG, this systematic variability becomes dominant over the rapidly fluctuating component, demonstrating that the model is in a regime dominated by the protostellar magnetic field. The cause of this variability is discussed further in detail below.

Figure 6 shows the distribution of accretion rates in 5-day blocks of simulation time for the different magnetic fields. The top sub-panels of the figure show the total accretion rate distribution across the entire time considered for analysis. For the 10 G, 500 G, and 1 kG cases, the distribution broadly follows a log-normal, although the center shifts slightly between these. In the 1 kG model, the distribution flattens at later times, impacted by the substantial drop in the accretion rate resulting from the protostar briefly entering a strong propeller phase. For the 2 kG case, there is a bimodal distribution peaking at higher and lower accretion values, with a tail towards high accretion rates. The lower accretion rate peak corresponds to the stable phase and the higher peak with the tail corresponds to the times when matter dumps from the truncation radius onto the protostellar surface with a subsequent decay in the accretion rate back to the stable phase.

Figure 6 also shows two different solid black points corresponding to the logarithmic average, logm˙delimited-⟨⟩subscript˙𝑚\langle\log\dot{m}_{*}\rangle⟨ roman_log over˙ start_ARG italic_m end_ARG start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT ⟩ (black circle), and linear time-window average, m˙bl=macc/Twindowsubscriptdelimited-⟨⟩˙𝑚blsubscript𝑚accsubscript𝑇window\langle\dot{m}\rangle_{\rm bl}=m_{\rm acc}/T_{\rm window}⟨ over˙ start_ARG italic_m end_ARG ⟩ start_POSTSUBSCRIPT roman_bl end_POSTSUBSCRIPT = italic_m start_POSTSUBSCRIPT roman_acc end_POSTSUBSCRIPT / italic_T start_POSTSUBSCRIPT roman_window end_POSTSUBSCRIPT (black diamond). The lower three magnetic field strength cases are generally similar indicating that accretion via the stable mode is dominant (e.g., not dominated by major bursts), although the 1 kG case starts to show some differences at later times. For the 2 kG case, there is substantially more mass being accreted during the bursts than the quiescent phase, and as such there is a substantial deviation of up to an order of magnitude between logm˙delimited-⟨⟩subscript˙𝑚\langle\log\dot{m}_{*}\rangle⟨ roman_log over˙ start_ARG italic_m end_ARG start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT ⟩ and m˙delimited-⟨⟩˙𝑚\langle\dot{m}\rangle⟨ over˙ start_ARG italic_m end_ARG ⟩.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 6: Distribution of the total accretion rate over \approx 5-day blocks for the 10 G (left) 500 G (middle left) and 1000 Gauss (middle right) and 2000 Gauss (right)protostars. The black points are the logarithmic average of the accretion rate and the diamonds are the time-averaged, e.g. m˙bl=macc/Twindowsubscriptdelimited-⟨⟩subscript˙𝑚blsubscript𝑚accsubscript𝑇window\langle\dot{m}_{*}\rangle_{\rm bl}=m_{\rm acc}/T_{\rm window}⟨ over˙ start_ARG italic_m end_ARG start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT roman_bl end_POSTSUBSCRIPT = italic_m start_POSTSUBSCRIPT roman_acc end_POSTSUBSCRIPT / italic_T start_POSTSUBSCRIPT roman_window end_POSTSUBSCRIPT, where maccsubscript𝑚accm_{\rm acc}italic_m start_POSTSUBSCRIPT roman_acc end_POSTSUBSCRIPT is the accretion mass during that block. Annotated is the block number (also denoted by color), and the time in the center of the block in days.

3.4 Accretion Bursts for Protostars with kG Magnetic fields

The 2 kG protostar model exhibits clearly defined accretion variability with a quasi-periodic signal defined by strong bursts (periods when the accretion rate grows rapidly). Further, the 1 kG protostar exhibits pseudo-periodic accretion bursts. In both of these models, the protostellar disk is truncated with gas flowing along magnetic flux tubes to the protostellar surface. In this subsection, we present, first, two different limiting regimes to estimate the critical magnetic field to truncate the disk, and then we discuss possible mechanisms to explain the accretion bursts.

3.4.1 Critical magnetic field strengths for disk truncation

The critical magnetic field at which the magnetic pressure truncates the disk can be described by balancing the ram pressure of the accreting gas (ρvr2𝜌superscriptsubscript𝑣𝑟2\rho v_{r}^{2}italic_ρ italic_v start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT) with the magnetic pressure of the protostellar dipole plus the disk magnetic field (B2/8πsuperscript𝐵28𝜋B^{2}/8\piitalic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 8 italic_π). The radius at which this occurs is called the truncation radius, RTsubscript𝑅𝑇R_{T}italic_R start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT, since the magnetic pressure halts the disk inflow, truncating the disk and producing a distinct separation between the disk and the accretion zone. We consider two different limiting regimes, i) the gas is infalling near free-fall velocities, and ii) the gas is advecting through the disk purely due to viscosity.

For the first limiting regime, we assume the gas is infalling at some fraction, fvsubscript𝑓𝑣f_{v}italic_f start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT, of the free-fall velocity (vff=2Gm/Rsubscript𝑣ff2𝐺subscript𝑚subscript𝑅v_{\rm ff}=\sqrt{2Gm_{*}/R_{*}}italic_v start_POSTSUBSCRIPT roman_ff end_POSTSUBSCRIPT = square-root start_ARG 2 italic_G italic_m start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT / italic_R start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT end_ARG) through some fraction, faccsubscript𝑓accf_{\rm acc}italic_f start_POSTSUBSCRIPT roman_acc end_POSTSUBSCRIPT, of a sphere around the protostar. The covering fraction is defined as facc=Aacc/Asubscript𝑓accsubscript𝐴accsubscript𝐴f_{\rm acc}=A_{\rm acc}/A_{*}italic_f start_POSTSUBSCRIPT roman_acc end_POSTSUBSCRIPT = italic_A start_POSTSUBSCRIPT roman_acc end_POSTSUBSCRIPT / italic_A start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT where Aaccsubscript𝐴accA_{\rm acc}italic_A start_POSTSUBSCRIPT roman_acc end_POSTSUBSCRIPT is the spherical surface area covered by the accretion flow and A=4πR2subscript𝐴4𝜋superscriptsubscript𝑅2A_{*}=4\pi R_{*}^{2}italic_A start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT = 4 italic_π italic_R start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT is the surface area of the protostar. We denote the magnitude of the total magnetic field as B=bB𝐵𝑏subscript𝐵B=bB_{*}italic_B = italic_b italic_B start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT, where b=|Bdisk+B|/|B|𝑏subscript@vecBdisksubscript@vecBsubscript@vecBb=|\@vec{B}_{\rm disk}+\@vec{B}_{*}|/|\@vec{B}_{*}|italic_b = | start_ID start_ARG italic_B end_ARG end_ID start_POSTSUBSCRIPT roman_disk end_POSTSUBSCRIPT + start_ID start_ARG italic_B end_ARG end_ID start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT | / | start_ID start_ARG italic_B end_ARG end_ID start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT |, to account for both the protostellar dipole and the disk magnetic field. In the case of anti-parallel fields, b𝑏bitalic_b may be less than 1. From this, we can write the critical magnetic field strength needed to make the radius at which the magnetic pressure can compete with the ram pressure outside the protostar radius as

(bB)M1kG(fv1.0)1/2(facc0.1)1/2(R3R)5/4×(m˙106Myr1)1/2(mM)1/4.subscript𝑏subscript𝐵𝑀1kGsuperscriptsubscript𝑓𝑣1.012superscriptsubscript𝑓acc0.112superscriptsubscript𝑅3subscript𝑅direct-product54superscriptsubscript˙𝑚superscript106subscript𝑀direct-productsuperscriptyr112superscriptsubscript𝑚subscript𝑀direct-product14(bB_{*})_{M}\approx 1\,{\rm kG}\left(\frac{f_{v}}{1.0}\right)^{1/2}\left(\frac% {f_{\rm acc}}{0.1}\right)^{-1/2}\left(\frac{R_{*}}{3\,R_{\odot}}\right)^{-5/4}% \\ \times\left(\frac{\dot{m}_{*}}{10^{-6}\,{M_{\odot}{\rm yr}^{-1}}}\right)^{1/2}% \left(\frac{m_{*}}{M_{\odot}}\right)^{1/4}.start_ROW start_CELL ( italic_b italic_B start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_M end_POSTSUBSCRIPT ≈ 1 roman_kG ( divide start_ARG italic_f start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT end_ARG start_ARG 1.0 end_ARG ) start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT ( divide start_ARG italic_f start_POSTSUBSCRIPT roman_acc end_POSTSUBSCRIPT end_ARG start_ARG 0.1 end_ARG ) start_POSTSUPERSCRIPT - 1 / 2 end_POSTSUPERSCRIPT ( divide start_ARG italic_R start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT end_ARG start_ARG 3 italic_R start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT - 5 / 4 end_POSTSUPERSCRIPT end_CELL end_ROW start_ROW start_CELL × ( divide start_ARG over˙ start_ARG italic_m end_ARG start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT end_ARG start_ARG 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT roman_yr start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT end_ARG ) start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT ( divide start_ARG italic_m start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT end_ARG start_ARG italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT 1 / 4 end_POSTSUPERSCRIPT . end_CELL end_ROW (25)

For the second limiting regime, we assume that the accretion onto the protostar is occurring purely via viscous flows through the disk. The geometric area of the accretion flow through the disk can be written as AfDRh𝐴subscript𝑓𝐷𝑅A\approx f_{D}Rhitalic_A ≈ italic_f start_POSTSUBSCRIPT italic_D end_POSTSUBSCRIPT italic_R italic_h, where fDsubscript𝑓𝐷f_{D}italic_f start_POSTSUBSCRIPT italic_D end_POSTSUBSCRIPT is the fraction of the disk cross-section through which gas is flowing, h=cs/ΩKsubscript𝑐𝑠subscriptΩ𝐾h=c_{s}/\Omega_{K}italic_h = italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT / roman_Ω start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT is the scale height under equilibrium, cssubscript𝑐𝑠c_{s}italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT is the sound speed and ΩKsubscriptΩ𝐾\Omega_{K}roman_Ω start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT is the Keplerian rotation frequency. The in-fall velocity through the disk, from mass conservation, is vr=m˙2πΣ(R)Rsubscript𝑣𝑟subscript˙𝑚2𝜋Σ𝑅𝑅v_{r}=\frac{\dot{m}_{*}}{2\pi\Sigma(R)R}italic_v start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT = divide start_ARG over˙ start_ARG italic_m end_ARG start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_π roman_Σ ( italic_R ) italic_R end_ARG where Σm˙/(3πηs)Σ˙𝑚3𝜋subscript𝜂𝑠\Sigma\approx\dot{m}/(3\pi\eta_{s})roman_Σ ≈ over˙ start_ARG italic_m end_ARG / ( 3 italic_π italic_η start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ). Using the approximation of a steady-state viscous accretion disk, determining the temperature as a balance of radiative cooling and viscous heating, and an α𝛼\alphaitalic_α-viscosity given by ηs=αcshsubscript𝜂𝑠𝛼subscript𝑐𝑠\eta_{s}=\alpha c_{s}hitalic_η start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = italic_α italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT italic_h, the critical magnetic field strength can be written as

(bB)α184Gauss(fV1.0)1/2(fD0.1)1/2(m˙106Myr1)9/16(R3R)19/16(α0.1)1/2(mM)1/16subscript𝑏subscript𝐵𝛼184Gausssuperscriptsubscript𝑓𝑉1.012superscriptsubscript𝑓𝐷0.112superscriptsubscript˙𝑚superscript106subscript𝑀direct-productsuperscriptyr1916superscriptsubscript𝑅3subscript𝑅direct-product1916superscript𝛼0.112superscriptsubscript𝑚subscript𝑀direct-product116(bB_{*})_{\alpha}\approx 184\,{\rm Gauss}\left(\frac{f_{V}}{1.0}\right)^{1/2}% \left(\frac{f_{D}}{0.1}\right)^{-1/2}\left(\frac{\dot{m}_{*}}{10^{-6}\,{M_{% \odot}{\rm yr}^{-1}}}\right)^{9/16}\\ \left(\frac{R_{*}}{3\,R_{\odot}}\right)^{-19/16}\left(\frac{\alpha}{0.1}\right% )^{1/2}\left(\frac{m_{*}}{M_{\odot}}\right)^{1/16}start_ROW start_CELL ( italic_b italic_B start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ≈ 184 roman_Gauss ( divide start_ARG italic_f start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT end_ARG start_ARG 1.0 end_ARG ) start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT ( divide start_ARG italic_f start_POSTSUBSCRIPT italic_D end_POSTSUBSCRIPT end_ARG start_ARG 0.1 end_ARG ) start_POSTSUPERSCRIPT - 1 / 2 end_POSTSUPERSCRIPT ( divide start_ARG over˙ start_ARG italic_m end_ARG start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT end_ARG start_ARG 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT roman_yr start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT end_ARG ) start_POSTSUPERSCRIPT 9 / 16 end_POSTSUPERSCRIPT end_CELL end_ROW start_ROW start_CELL ( divide start_ARG italic_R start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT end_ARG start_ARG 3 italic_R start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT - 19 / 16 end_POSTSUPERSCRIPT ( divide start_ARG italic_α end_ARG start_ARG 0.1 end_ARG ) start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT ( divide start_ARG italic_m start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT end_ARG start_ARG italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT 1 / 16 end_POSTSUPERSCRIPT end_CELL end_ROW (26)

As shown, (bB)αsubscript𝑏subscript𝐵𝛼(bB_{*})_{\alpha}( italic_b italic_B start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT is typically lower than (bB)Msubscript𝑏subscript𝐵𝑀(bB_{*})_{M}( italic_b italic_B start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_M end_POSTSUBSCRIPT since the free fall velocity is greater than the disk accretion velocity. In our simulations, we find that the gas is flowing at sub-free-fall speeds, but faster than the viscous speeds, agreeing with the disk becoming truncated above B>500subscript𝐵500B_{*}>500italic_B start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT > 500 G. We note that the above equations are approximate since they follow the assumptions that the disk is geometrically thin and the mass and angular momentum transport are in a steady-state equilibrium. However, they provide a rough estimate of this critical field strength.

In the 1 kG and 2 kG simulations, gas builds up at the truncation radius in a stable knot until its mass penetrates the magnetosphere as a filament and funnels onto the protostar, traveling along nearby magnetic fields to the surface. We find that the gas does not hit exactly at the equator and is instead guided by field lines that connect to the surface at low altitudes.

3.4.2 Accretion Bursting Mechanisms in the kG simulations

We consider two plausible mechanisms to understand the bursting demonstrated in Figure 5. For the first, we build a simple analytic prescription in which gas accreting through the disk builds up at the truncation radius until the momentum of the gas can overcome the magnetic pressure, funneling all the gas onto the surface in a burst. We compare the predictions from this model with one of the major bursts from the 2 kG simulation. The density of the gas being built up at the truncation radius by

ρms=m˙×tbfA×2πRTΔr2,subscript𝜌mssubscript˙𝑚subscript𝑡𝑏subscript𝑓𝐴2𝜋subscript𝑅TΔsuperscript𝑟2\rho_{\rm ms}=\frac{\dot{m}_{*}\times t_{b}}{f_{A}\times 2\pi R_{\rm T}\Delta r% ^{2}},italic_ρ start_POSTSUBSCRIPT roman_ms end_POSTSUBSCRIPT = divide start_ARG over˙ start_ARG italic_m end_ARG start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT × italic_t start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT end_ARG start_ARG italic_f start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT × 2 italic_π italic_R start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT roman_Δ italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG , (27)

where fAsubscript𝑓𝐴f_{A}italic_f start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT is the filling fraction of the torus around the protostar (in our simulations, fA=1subscript𝑓𝐴1f_{A}=1italic_f start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT = 1 by definition since we do not include the azimuthal dimension), ΔrΔ𝑟\Delta rroman_Δ italic_r is the radial width of the torus, and tbsubscript𝑡𝑏t_{b}italic_t start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT is the time since the last accretion event. Assuming mass conservation outside of the truncation radius, i.e., that the radial velocity of the gas flowing against the stable knot of gas is vr=fvvffsubscript𝑣𝑟subscript𝑓𝑣subscript𝑣ffv_{r}=f_{v}v_{\rm ff}italic_v start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT = italic_f start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT roman_ff end_POSTSUBSCRIPT, where fvsubscript𝑓𝑣f_{v}italic_f start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT is the fraction of gas in-fall speed to the free-fall speed, the time between bursts is

tbB2fA(RTΔr)28GfV2mm˙,subscript𝑡𝑏superscript𝐵2subscript𝑓𝐴superscriptsubscript𝑅TΔ𝑟28𝐺superscriptsubscript𝑓𝑉2subscript𝑚˙𝑚t_{b}\approx\frac{B^{2}f_{A}(R_{\rm T}\Delta r)^{2}}{8Gf_{V}^{2}m_{*}\dot{m}},italic_t start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ≈ divide start_ARG italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_f start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ( italic_R start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT roman_Δ italic_r ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 8 italic_G italic_f start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_m start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT over˙ start_ARG italic_m end_ARG end_ARG , (28)

where B𝐵Bitalic_B is the local magnetic field magnitude. For the burst around day 45 of the simulation, we find that the velocity of the gas in the disk midplane is not free-falling but moving at approximately 10% free-fall, consistent with the average accretion rate. The width of the torus can be estimated by assuming the gas falling in hits an impenetrable wall and oscillates around that spot in a stable Keplerian orbit, such that width is balanced by the length scale corresponding to a sound wave traveling over a full orbit. The width of the torus then follows

Δrcs×2πΩK(RT),Δ𝑟subscript𝑐𝑠2𝜋subscriptΩ𝐾subscript𝑅𝑇\Delta r\approx c_{s}\times\frac{2\pi}{\Omega_{K}(R_{T})},roman_Δ italic_r ≈ italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT × divide start_ARG 2 italic_π end_ARG start_ARG roman_Ω start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT ( italic_R start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT ) end_ARG , (29)

where cssubscript𝑐𝑠c_{s}italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT is the thermal sound speed and ΩK(RT)subscriptΩ𝐾subscript𝑅𝑇\Omega_{K}(R_{T})roman_Ω start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT ( italic_R start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT ) is the Keplerian angular speed at the truncation radius. Note that under this assumption, the width of the torus scales similarly to the scale height of the disk at the truncation radius. For a truncation radius, RT0.02subscript𝑅𝑇0.02R_{T}\approx 0.02italic_R start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT ≈ 0.02 AU and a gas temperature 104absentsuperscript104\approx 10^{4}≈ 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT Kelvin, the torus width is approximately 0.005 AU, in rough agreement with the empirically found width around the bursts. Using the simulation results around this burst, and m˙=106Msubscript˙𝑚superscript106subscript𝑀direct-product\dot{m}_{*}=10^{-6}M_{\odot}over˙ start_ARG italic_m end_ARG start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT yr-1, the above expression gives tb10subscript𝑡��10t_{b}\approx 10italic_t start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ≈ 10 days, in qualitative agreement with the simulation results.

However, one would expect that under steady-state background accretion through the disk, the above mechanism would produce a clear signature of periodicity in the accretion rate. Figure 7 shows the Lomb-Scargle periodograms for the 1 kG and 2 kG models. We find that there are no clear dominating periodic signals, with power being distributed across a range of frequency peaks. The 2 kG model shows the strongest signal at roughly 3.5 and 7 days, although there is a broad distribution of power at signals from approximately 12 hours to 10 days.

Refer to caption
Refer to caption
Figure 7: Lomb-scargle periodogram for the 1 kG (top) and 2 kG (bottom) models. The vertical dotted line shows the stellar rotation period, and the dashed line shows the Nyquist frequency for the duration of the simulation. The red points highlight peaks with a false alarm probability <1%absentpercent1<1\%< 1 %.

Due to the lack of clear periodicity, we also consider the magnetic Rayleigh-Taylor instability, also called the “interchange instability” (Kruskal & Schwarzschild, 1954; Newcomb, 1961). This instability has been found numerically to exist in magnetospheric accretion in T-Tauri-like protostars (Kulkarni & Romanova, 2008; Blinova et al., 2016; Zhu et al., 2024). We utilize the “fastness” parameter (e.g. Ghosh, 2007; Romanova et al., 2018), defined here as

ωs=ΩΩ(RT),subscript𝜔𝑠subscriptΩΩsubscript𝑅𝑇\omega_{s}=\frac{\Omega_{*}}{\Omega(R_{T})},italic_ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = divide start_ARG roman_Ω start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT end_ARG start_ARG roman_Ω ( italic_R start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT ) end_ARG , (30)

where Ω(RT)=vϕ/RΩsubscript𝑅𝑇subscript𝑣italic-ϕ𝑅\Omega(R_{T})=v_{\phi}/Rroman_Ω ( italic_R start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT ) = italic_v start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT / italic_R is the rotation rate of the disk at the truncation radius. We consider the general rotation rate rather than the Keplerian rate since, although most of the disk vϕ=vkepsubscript𝑣italic-ϕsubscript𝑣kepv_{\phi}=v_{\rm kep}italic_v start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT = italic_v start_POSTSUBSCRIPT roman_kep end_POSTSUBSCRIPT, in the inner region there can be departures from Keplerian rotation. We empirically define the truncation radius as the radius at which the total magnetic pressure is balanced by the total ram pressure and thermal pressure (Romanova et al., 2018), i.e.

B28π=ρv2+p,superscript𝐵28𝜋𝜌superscript𝑣2𝑝\frac{B^{2}}{8\pi}=\rho v^{2}+p,divide start_ARG italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 8 italic_π end_ARG = italic_ρ italic_v start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_p , (31)

at R=RT𝑅subscript𝑅𝑇R=R_{T}italic_R = italic_R start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT. We use the general expression since the protostar is rotating and thus at any point could be in a propeller regime. Outside of the propeller regime, ρv2𝜌superscript𝑣2\rho v^{2}italic_ρ italic_v start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT approaches ρvr2𝜌superscriptsubscript𝑣𝑟2\rho v_{r}^{2}italic_ρ italic_v start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT which is often used ram pressure term to balance the magnetic pressure (see above subsection, and review by Hartmann et al. (2016)). Blinova et al. (2016) found from a suite of simulations that the gas at the truncation radius could become unstable to the interchange instability when ωs<0.6subscript𝜔𝑠0.6\omega_{s}<0.6italic_ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT < 0.6. While the simulations presented herein include a more massive disk than Blinova et al. (2016), we consider this definition to roughly estimate when the truncation region may undergo this instability.

Refer to caption
Refer to caption
Figure 8: The fastness parameter, ωssubscript𝜔𝑠\omega_{s}italic_ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, as a function of time for the 1 kG (top) and 2 kG (bottom) protostar models. The ‘
‘ hatched region indicates where ωs<1subscript𝜔𝑠1\omega_{s}<1italic_ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT < 1 and the ‘X‘ hatched region indicates where ωs<0.6subscript𝜔𝑠0.6\omega_{s}<0.6italic_ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT < 0.6, and thus prone to the interchange instability. The cyan line indicates the daily rolling average.

Figure 8 shows the ωssubscript𝜔𝑠\omega_{s}italic_ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT for the 1 kG and 2 kG models. For the 1 kG model, the fastness parameter oscillates significantly between ωs>>1much-greater-thansubscript𝜔𝑠1\omega_{s}>>1italic_ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT > > 1 and ωs1subscript𝜔𝑠1\omega_{s}\approx 1italic_ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ≈ 1. While this indicates the protostar is close to the propeller regime, the inner truncation radius does not appear to be strongly in the regime of the instability. Further, the daily rolling average oscillates around ωs1subscript𝜔𝑠1\omega_{s}\approx 1italic_ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ≈ 1. However, it is noteworthy that the period of ωs>>1much-greater-thansubscript𝜔𝑠1\omega_{s}>>1italic_ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT > > 1 around day 50 is followed by a significant decrease in the accretion rate for a week. The decrease in the accretion is likely thus due to the protostar being momentarily in the strong propellor regime and the subsequent winds decreasing the accretion rate.

For the 2 kG case, there is much less rapid stochasticity in ωssubscript𝜔𝑠\omega_{s}italic_ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, with a significant amount of time showing ωs0.6subscript𝜔𝑠0.6\omega_{s}\approx 0.6italic_ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ≈ 0.6. Therefore, the pseudo-periodic variability in the accretion rate in this model may be readily explained by the interchange instability. Similarly to the 1 kG case, the only time window in which the rolling average ωssubscript𝜔𝑠\omega_{s}italic_ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT greatly exceeds 1 (around day 25) is followed by a significant reduction in the accretion rate. Our results suggest that accretion burst signatures are directly related to the protostellar magnetic field and the star-disk interaction via magnetic coupling. In both cases, the protostar is not entirely within the strong propeller regime, but instead randomly enters it. When this occurs, the magnetosphere expands, suddenly collapsing the gas flow and decreasing the amount of gas in the accretion region, thereby decreasing the accretion rate.

3.5 Accretion Luminosity Variability

The luminosity of protostars is observed to vary significantly, with both short and long periods found in Class 0/I protostars (see recent review by Fischer et al., 2023). While the HADES simulations do not trace the monthly or yearly periods, the simulations show significant variability on shorter timescales. The bolometric luminosity here is defined as

Lbol=L+Lacc,subscript𝐿bolsubscript𝐿subscript𝐿accL_{\rm bol}=L_{*}+L_{\rm acc},italic_L start_POSTSUBSCRIPT roman_bol end_POSTSUBSCRIPT = italic_L start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT + italic_L start_POSTSUBSCRIPT roman_acc end_POSTSUBSCRIPT , (32)

where L=3.5subscript𝐿3.5L_{*}=3.5italic_L start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT = 3.5 L is the intrinsic luminosity of the protostar, calculated using fits from Fischer et al. (2017) of the stellar tracks from Siess et al. (2000), and Laccsubscript𝐿accL_{\rm acc}italic_L start_POSTSUBSCRIPT roman_acc end_POSTSUBSCRIPT is the accretion luminosity, defined as

Lacc=Gmm˙R.subscript𝐿acc𝐺subscript𝑚subscript˙𝑚subscript𝑅L_{\rm acc}=\frac{Gm_{*}\dot{m}_{*}}{R_{*}}.italic_L start_POSTSUBSCRIPT roman_acc end_POSTSUBSCRIPT = divide start_ARG italic_G italic_m start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT over˙ start_ARG italic_m end_ARG start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT end_ARG start_ARG italic_R start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT end_ARG . (33)

We do not consider here individual bands or line emission, nor the reprocessing of the emission (which would reduce the emission as some of the radiation is absorbed by the star). Figure 9 shows the light curves for each of the four different models.

Refer to caption
Figure 9: Bolometric accretion luminosity versus time for the different protostellar magnetic field models, annotated on the plots. The cyan band denotes the starting region for the sampled light curves. The black dashed-dot line is the daily rolling average.
Refer to caption
Figure 10: Sampling points of the light curve mimicking the YSOVAR survey.

The instantaneous light curves exhibit significant differences between the models. In particular, the 10 G and 500 G models have less daily variation in their light curves but have significant variability on rapid time scales of an hour or less. As the magnetic field increases, the hourly variability decreases. For the 1 kG model, the daily average light curve is remarkably flat except for the period after 55 days when the protostar accretion significantly dips. The 2 kG shows a clear systematic variability with much less hourly variability. The emission at early times is dominated by the intrinsic luminosity, and at later times the light curve is dominated by peaks from the accretion bursts. Observationally, unless the protostar is observed multiple times per hour, the 10 G, 500 G, and 1 kG models would exhibit relatively flat light curves, except for the 1 kG model entering the strong propeller regime. The clear variability in the 2 kG exhibits a nearly exponential decrease which may be a clear signature of a highly magnetized protostar accreting via instbility.

Observations of sources in Serpens South by YSOVAR (Wolk et al., 2018) exhibit a range of light curves, including some consistent with those here. In particular, they find line curves with rapid but small-scale variability (e.g. SSTYSV J183000.56-020211.4, SSTYSV J183005.33-020111.8, SSTYSV J183005.81-020205.8) and others with larger broader variability (e.g. SSTYSV J183001.26-020148.3). Their object SSTYSV J183006.13-020108.0 shows a significant dip in luminosity, which they calculate has a period of approximately 35 days. However, this matches their sampling period, so it may be that the dip is not periodic variability, but instead from catching a protostar dipping due to entering the strong propeller regime and leaving it.

Observational campaigns to study protostellar variability are restricted in their sampling rate. One of the most robust surveys is YSOVAR (Rebull et al., 2014; Wolk et al., 2015, 2018), which utilized a logarithmic-spaced sampling strategy to trace both shorter and longer periods. We mimic the sampling strategy of YSOVAR, as shown in Figure 10, to produce synthetic light curve samples of our models. To produce a light curve for an “average protostar” similar to our models, we compute a synthetic “population average” YSOVAR light curve. We choose random starting points for the “observation” of the light curve, uniformly distributed in the blue bands shown in Figure 9. Starting from this initial time, we generate a light curve sampled following the Figure 10 sampling strategy to generate a population of light curves. Figure 11 shows these samples along with the logarithmic average and standard deviations of this population. The figure shows there is substantial scatter in the individual samples, with some sampled light curves showing very high luminosities. However, the average behavior is substantially milder. Figure 12 highlights the logarithmically averaged light curves and the standard deviations at each sample point. The variabilities, rather than orders of magnitude, are only factors of a few. Defining the magnitude variation as ΔM=log2.5σ(Lacc,samp)Δ𝑀subscript2.5𝜎delimited-⟨⟩subscript𝐿accsamp\Delta M=\log_{2.5}\sigma(\langle L_{\rm acc,samp}\rangle)roman_Δ italic_M = roman_log start_POSTSUBSCRIPT 2.5 end_POSTSUBSCRIPT italic_σ ( ⟨ italic_L start_POSTSUBSCRIPT roman_acc , roman_samp end_POSTSUBSCRIPT ⟩ ), we find ΔMΔ𝑀absent\Delta M\approxroman_Δ italic_M ≈ 0.3, 0.39, 0.12 and 0.78 for the 10 G, 500 G, 1 kG, and 2 kG models, respectively. These are in agreement with the variability measurements in Wolk et al. (2018). Much of the accretion variability is similar in this regard, however, for the 1 kG case it is smaller due to the more ordered fields which are not strong enough yet to induce the large variability seen in the 2 kG. In general, we find the 500 G model can become the brightest due to the enhanced accretion caused by the outflow cavity focusing the accretion flow, while the strongly magnetized sources are the dimmest.

Unfortunately, the results of the sampling mean that even with the incredible cadence of YSOVAR, constraining the hourly variability will be difficult. The order of magnitude variability in Figure 9 is substantially reduced, on average, at the cadence sampling of YSOVAR. The YSOVAR sampling strategy is highly useful if one wishes to trace both shorter and long-term evolution with minimal samples. Instead, a campaign using a rapid (<1absent1<1< 1 hour) linearly-spaced sampling strategy over several nights would be necessary to constrain any rapid variability.

Refer to caption
Figure 11: Synthetic light curve samples (black points) with the average and standard deviations given by the colored points.
Refer to caption
Figure 12: Average light curves from the samples in Fig. 11 with the standard deviation.

3.6 Accretion Filling Fraction

The filling fraction of the accretion onto the protostar is of key importance for models of star formation and protostar evolution. We computed the filling fraction of accretion for the top 10, 25, 50, 75, and 90% of the accreted mass as a function of time. To compute the filling fraction, for each snapshot, we sort the cells along the protostellar surface according to the mass flux, from highest flux to lowest. Marching down from the cells with the most mass being accreted to the least enables a better constraint on the effective area where a given percentage of the mass is being accreted. Using the flux and the cell surface area, we calculate the cumulative mass accretion fraction. Then, for the different top mass-accretion percentages we compute the total area of the cells marching down in mass flux which encompasses that percentage of the total accretion. Figure 13 shows an example of a single snapshot from the 2 kG model, plotting the sorted mass flux versus the index, along with the cumulative mass accretion, area, and the cell’s accretion rate. For this snapshot, Figure 14 shows the accumulated mass fraction (marching up from the lowest) versus the area fraction with lines denoting the 10, 25, 50, 75, and 90% for this snapshot example. The figure shows that the vast majority of the mass is accreted over a minority of the surface.

Refer to caption
Figure 13: Example of how the filling fraction is computed for a single snapshot. The figure shows the sorted mass flux at the surface (red), accretion rate (green), cumulative mass accretion (black), and cumulative area (blue) versus the sorted cell index. The horizontal lines show the index at which the top 10%, 25%, 50%, 75%, and 90% of the mass is accreted.
Refer to caption
Figure 14: The accumulated mass fraction percentage versus the cumulative area of the protostar surface for the snapshot in Figure 13. The horizontal lines show the index at which the top 10%, 25%, 50%, 75%, and 90% of the mass is accreted.

Figure 15 shows the filling fraction of accretion versus time for the 2 kG protostar, plotted using the daily rolling average for ease of visualization and interpretation. Table 1 shows a summary of the accretion-rate-weighted average filling fraction over the simulation period for the four different models. We find that the top 25% of the mass is accreted in about 1% of the surface area. The majority of the mass, at 50%, is accreted over 2-4% of the surface. Finally, of most importance, 90% of the accreted mass is from mass flux encompassing between 7 - 20% of the surface. In all models, the last 10% of the mass is in general accreted more diffusely over the surface (see e.g. Figure 14, where the last 10% of the mass is accreted over 70% of the area). In general, we find that these filling fractions are roughly equal with time, however, they can increase by factors of 2-3 during accretion bursts, as seen in Figure 15. Depending on the protostellar magnetic field assumed, a filling fraction of protostar accretion between 10-20% is reasonable for protostars accreting under these modes.

Table 1: Filling fraction (percent) of the accretion flow onto the protostar according to the top X% of the mass accretion.
B 10% 25% 50% 75% 90%
10 G 0.4 1.0 2.5 6.6 15.9
500 G 0.6 1.5 4.2 10.2 19.7
1000 G 0.5 1.2 3.1 6.8 12.1
2000 G 0.4 0.8 1.9 4.0 7.4
Refer to caption
Figure 15: Top: The accretion rate versus time for the 2 kG protostar, using the daily-rolling average. The daily-rolling-average filling fraction of the accretion flow for the different top mass accretion percentages.

4 Discussion

After analyzing these simulations, we find that there are three distinct regimes of the accretion: i) boundary layer (10 G, 500 G), ii) turbulent magnetospheric (1 kG), and iii) magnetospheric accompanied by the interchange instability (2 kG).

Figure 16 visualizes the importance of the magnetic field via the ratio of the poloidal pressures, (ρvp2+p)/[(Br2+Bθ2)/2]𝜌superscriptsubscript𝑣𝑝2𝑝delimited-[]superscriptsubscript𝐵𝑟2superscriptsubscript𝐵𝜃22(\rho v_{p}^{2}+p)/[(B_{r}^{2}+B_{\theta}^{2})/2]( italic_ρ italic_v start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_p ) / [ ( italic_B start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_B start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) / 2 ], in the inner region of the disk. The figure shows the average ratio from days 50-51 of the simulation and the lines denote the direction of the poloidal magnetic field. The 10 G model shows the gas fully impacts the protostar surface with only a marginally magnetically dominated pole. The 500 G model immediately demonstrates the impact of the magnetically dominated cavity creating a pressure that inhibits gas flow to the poles, focusing the accretion to the equator. The kG-strength models show clearly magnetically truncated disks. In all cases, towards outer radii (0.2 AU), the magnetic fields transition to vertical. In the inner disk, there is a substantial difference between the models, especially due to the impact of the Hall effect for the cases with strong magnetospheres. Inside the disk, there is a strong correlation between the morphology of the field lines, the presence of magnetic islands, and the ratio of the pressures.

Refer to caption
Figure 16: The ratio of the poloidal ram pressure plus thermal pressure to the poloidal magnetic pressure within the inner 0.2 AU of the disk. The black arrows indicate the direction of the poloidal component of the magnetic field. The ratio is averaged over 1 day starting from day 50. The panels are in ascending order of protostar magnetic field, annotated on the top, of 10 G, 500 G, 1 kG, and 2 kG from top to bottom.

Takasao et al. (2018) simulated an accretion disk around an unmagnetized central object, with the T-Tauri scaling providing a nominal accretion rate of m˙108Msubscript˙𝑚superscript108subscript𝑀direct-product\dot{m}_{*}\approx 10^{-8}M_{\odot}over˙ start_ARG italic_m end_ARG start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT ≈ 10 start_POSTSUPERSCRIPT - 8 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT yr-1, a factor of 100 less than in this study. They find ordered boundary layer accretion with an accretion component due to high-latitude funnel wall flows. This is in broad agreement with our 10 G simulation, where we find that the protostellar disk directly impacts the protostar surface, with two high latitude, nearly constant, inflows due to failed disk winds cycling back down to the disk surface and sliding along the surface. However, the accretion variability we see is significantly higher than theirs (see their Fig. 19), which may be due to the difference in the disk surface density and the overall accretion rate magnitude difference. Mass ejections from the protostar surface can disrupt the accretion momentarily, leading to dips in the averaged accretion rate. Takasao et al. (2018), similarly to these simulations, find that the midplane component, especially at later times, becomes complemented by higher latitude flows. However, the high latitude flows in the results presented herein occur at θ>30𝜃superscript30\theta>30^{\circ}italic_θ > 30 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT, rather than 15 they find. While the initial density distributions between the simulations differ, both simulations use hour-glass morphologies for the disk magnetic field.

Our simulations also tell us why the boundary layer accretion in the 500 G protostar is ordered and enhanced. The protostellar magnetic field induces a strongly magnetized, ordered bipolar outflow. This magnetized cavity produces a pressure pushing the inner accretion flow down, constraining and focusing it towards the protostar’s equator. While the presence of the protostellar magnetic field does not truncate the disk and create well-defined accretion columns, the resulting outflow orders the accretion flow away from the protostellar poles, funneling it towards the equation, boosting the accretion rate compared to the 10 G protostar. As a protostellar magnetic field is increased, the accretion flow will become more ordered, transitioning from a turbulent boundary layer mode to a more ordered boundary layer mode, eventually to the magnetospheric mode at kG scales with bursts.

5 Conclusions

We have presented the initial suite of high-resolution simulations of accretion disks around embedded, actively accreting protostars. We model protostars with four different protostellar magnetic fields, 10 G, 500 G, 1 kG, and 2 kG. These simulations resolve the inner accretion disk with a maximal resolution of \approx 10-4 AU. Our simulations show that

  1. 1.

    The protostellar magnetic field plays a crucial role in determining the underlying accretion mechanism. We find that for solar-mass protostars accreting at rates exceeding 107superscript10710^{-7}10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT M yr-1, protostars must generate a magnetic field exceeding 1 kG to truncate the disk and produce magnetospheric accretion. For weaker fields, we find that boundary layer accretion dominates.

  2. 2.

    For poorly magnetized protostars (10 G), the accretion falls across the entire surface of the protostar. The gas impacts both at the equator and via two high-latitude streamers which are the result of failed winds falling back in.

  3. 3.

    For marginally magnetized protostars (500 G), in which the disk is not truncated, gas accretes onto the protostar via boundary layer accretion. However, the presence of a magnetically-driven bipolar outflow focuses the accretion flow towards the protostar’s equator, enhancing and regulating the accretion rate.

  4. 4.

    For kG-strength fields, we find there is substantial systematic variability, with the 2 kG model exhibiting strong bursts of accretion. The most likely explanation for the accretion bursts in the 2 kG is the interchange instability occurring at the truncation radius.

  5. 5.

    Our 1 kG and 2 kG simulations have one-sided outflows due to the symmetry of the protostellar magnetic dipole being broken by the disk magnetic field.

  6. 6.

    In our simulations, the protostars accrete the vast majority of the mass across 10-20% of their surface area. While there is some variance in the filling fraction between the protostellar magnetic fields, for protostar accretion modeling and sub-grid models we’d recommend a filling fraction of 10-20% for protostars at these stages of evolution.

We investigated the variability in the accretion luminosity for the four different models. The instantaneous light curve found in our simulation varies by factors of a few to orders of magnitude depending on the magnetic field. For the 10 G, 500 G, and 1 kG models, the variability is generally occurring at hourly timescales while the B=2subscript𝐵2B_{*}=2italic_B start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT = 2 kG model exhibits significant variability of over an order of magnitude over days due to accretion bursts. The bolometric accretion luminosity computed this way does not account for radiation transfer effects. Observations of variability do not sample uniformly for sub-hourly timescales continuously but instead employ a range of strategies such as the logarithmic sampling of YSOVAR. Therefore, we produce averaged synthetic light curves using a cadence to mimic YSOVAR (Rebull et al., 2014). We find that there is far less variability in these average light curves, with bolometric magnitude variances less than one, roughly consistent with observations (Wolk et al., 2018).

These simulations motivate further observations of magnetic fields towards Class I protostars. The recent survey from Flores et al. (2024) measured the magnetic fields for a range of Class I sources using high-resolution K-band spectroscopy. Their sample exhibits magnetic fields ranging from 500 G to nearly 4 kG, with the majority of the sample having fields between 1 - 2 kG. While the survey is not sensitive to unmagnetized protostars, a key takeaway is that the magnetic field does exhibit a broad spread, and as such our simulation results across a range of protostellar magnetic fields can aid in interpreting accretion observations of these earlier protostars. Future work will investigate synthetic Hydrogen recombination line emission, a broader parameter space for the protostar rotation, accretion rate, disk magnetic field morphologies, and the physics of outflow launching.

Acknowledgements.
BALG is supported by the Chalmers Initiative on Cosmic Origins as a Cosmic Origins Postdoctoral Fellow. JCT acknowledges support from ERC Advanced Grant 788829 (MSTAR). RK acknowledges financial support via the Heisenberg Research Grant funded by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) under grant no. KU 2849/9, project no. 445783058. The computations were enabled by resources provided by the National Academic Infrastructure for Supercomputing in Sweden (NAISS) at Dardel/PDC partially funded by the Swedish Research Council through grant agreement no. 2022-06725 through the allocations SNIC 2022/5-654 and NAISS 2023/23-341 and resources provided by Chalmers e-Commons at Chalmers. This work made use of the following software packages: matplotlib (Hunter, 2007), numpy (Harris et al., 2020) and python (Van Rossum & Drake, 2009). Parts of the results in this work make use of the colormaps in the CMasher (van der Velden, 2020; van der Velden et al., 2024) and cmocean (Thyng et al., 2016). Software citation information aggregated using The Software Citation Station (Wagg & Broekgaarden, 2024, 2024).

References

  • Ahmad et al. (2024) Ahmad, A., González, M., Hennebelle, P., & Commerçon, B. 2024, A&A, 687, A90
  • Avison et al. (2021) Avison, A., Fuller, G. A., Peretto, N., et al. 2021, A&A, 645, A142
  • Balbus & Hawley (2000) Balbus, S. A. & Hawley, J. F. 2000, Space Sci. Rev., 92, 39
  • Balbus & Terquem (2001) Balbus, S. A. & Terquem, C. 2001, ApJ, 552, 235
  • Bate et al. (2014) Bate, M. R., Tricco, T. S., & Price, D. J. 2014, MNRAS, 437, 77
  • Béthune et al. (2017) Béthune, W., Lesur, G., & Ferreira, J. 2017, A&A, 600, A75
  • Bhandare et al. (2020) Bhandare, A., Kuiper, R., Henning, T., et al. 2020, A&A, 638, A86
  • Bleuler & Teyssier (2014) Bleuler, A. & Teyssier, R. 2014, MNRAS, 445, 4015
  • Blinova et al. (2016) Blinova, A. A., Romanova, M. M., & Lovelace, R. V. E. 2016, MNRAS, 459, 2354
  • Bonnell et al. (2001) Bonnell, I. A., Bate, M. R., Clarke, C. J., & Pringle, J. E. 2001, MNRAS, 323, 785
  • Calvet & Gullbring (1998) Calvet, N. & Gullbring, E. 1998, ApJ, 509, 802
  • Carroll et al. (2012) Carroll, T. A., Strassmeier, K. G., Rice, J. B., & Künstler, A. 2012, A&A, 548, A95
  • Chiang & Goldreich (1997) Chiang, E. I. & Goldreich, P. 1997, ApJ, 490, 368
  • Commerçon et al. (2022) Commerçon, B., González, M., Mignon-Risse, R., Hennebelle, P., & Vaytet, N. 2022, A&A, 658, A52
  • Cunningham et al. (2011) Cunningham, A. J., Klein, R. I., Krumholz, M. R., & McKee, C. F. 2011, ApJ, 740, 107
  • Cunningham et al. (2018) Cunningham, A. J., Krumholz, M. R., McKee, C. F., & Klein, R. I. 2018, MNRAS, 476, 771
  • Dale et al. (2012) Dale, J. E., Ercolano, B., & Bonnell, I. A. 2012, MNRAS, 424, 377
  • Dalgarno et al. (1999) Dalgarno, A., Yan, M., & Liu, W. 1999, ApJS, 125, 237
  • Dere (2013) Dere, K. 2013, ChiantiPy: Python package for the CHIANTI atomic database, Astrophysics Source Code Library, record ascl:1308.017
  • Dere et al. (2019) Dere, K. P., Del Zanna, G., Young, P. R., Landi, E., & Sutherland, R. S. 2019, ApJS, 241, 22
  • Donati et al. (2007) Donati, J. F., Jardine, M. M., Gregory, S. G., et al. 2007, MNRAS, 380, 1297
  • Donati et al. (2008) Donati, J. F., Jardine, M. M., Gregory, S. G., et al. 2008, MNRAS, 386, 1234
  • Federman et al. (2024) Federman, S. A., Megeath, S. T., Rubinstein, A. E., et al. 2024, ApJ, 966, 41
  • Feigelson et al. (2007) Feigelson, E., Townsley, L., Güdel, M., & Stassun, K. 2007, in Protostars and Planets V, ed. B. Reipurth, D. Jewitt, & K. Keil, 313
  • Feigelson & Montmerle (1999) Feigelson, E. D. & Montmerle, T. 1999, ARA&A, 37, 363
  • Fiorellino et al. (2023) Fiorellino, E., Tychoniec, Ł., Cruz-Sáenz de Miera, F., et al. 2023, ApJ, 944, 135
  • Fiorellino et al. (2022) Fiorellino, E., Tychoniec, Ł., Manara, C. F., et al. 2022, ApJ, 937, L9
  • Fischer et al. (2023) Fischer, W. J., Hillenbrand, L. A., Herczeg, G. J., et al. 2023, in Astronomical Society of the Pacific Conference Series, Vol. 534, Protostars and Planets VII, ed. S. Inutsuka, Y. Aikawa, T. Muto, K. Tomida, & M. Tamura, 355
  • Fischer et al. (2017) Fischer, W. J., Megeath, S. T., Furlan, E., et al. 2017, ApJ, 840, 69
  • Flores et al. (2024) Flores, C., Connelley, M. S., Reipurth, B., Boogert, A., & Doppmann, G. 2024, arXiv e-prints, arXiv:2405.12451
  • Fromang et al. (2002) Fromang, S., Terquem, C., & Balbus, S. A. 2002, MNRAS, 329, 18
  • Gaches et al. (2023) Gaches, B. A. L., Walch, S., Wünsch, R., & Mackey, J. 2023, MNRAS, 522, 4674
  • Geroux et al. (2016) Geroux, C., Baraffe, I., Viallet, M., et al. 2016, A&A, 588, A85
  • Getman et al. (2022) Getman, K. V., Feigelson, E. D., Garmire, G. P., et al. 2022, ApJ, 935, 43
  • Ghosh (2007) Ghosh, P. 2007, Rotation and Accretion Powered Pulsars, Vol. 10
  • Gravity Collaboration et al. (2023a) Gravity Collaboration, Soulain, A., Perraut, K., et al. 2023a, A&A, 674, A203
  • Gravity Collaboration et al. (2023b) Gravity Collaboration, Wojtczak, J. A., Labadie, L., et al. 2023b, A&A, 669, A59
  • Gregory et al. (2008) Gregory, S. G., Matt, S. P., Donati, J. F., & Jardine, M. 2008, MNRAS, 389, 1839
  • Grudić et al. (2021) Grudić, M. Y., Guszejnov, D., Hopkins, P. F., Offner, S. S. R., & Faucher-Giguère, C.-A. 2021, MNRAS, 506, 2199
  • Harris et al. (2020) Harris, C. R., Millman, K. J., van der Walt, S. J., et al. 2020, Nature, 585, 357
  • Hartmann et al. (2016) Hartmann, L., Herczeg, G., & Calvet, N. 2016, ARA&A, 54, 135
  • Herbst et al. (2007) Herbst, W., Eislöffel, J., Mundt, R., & Scholz, A. 2007, in Protostars and Planets V, ed. B. Reipurth, D. Jewitt, & K. Keil, 297
  • Hunter (2007) Hunter, J. D. 2007, Computing in Science & Engineering, 9, 90
  • Ireland et al. (2021) Ireland, L. G., Zanni, C., Matt, S. P., & Pantolmos, G. 2021, ApJ, 906, 4
  • Kenyon et al. (1993) Kenyon, S. J., Calvet, N., & Hartmann, L. 1993, ApJ, 414, 676
  • Kley & Lin (1996) Kley, W. & Lin, D. N. C. 1996, ApJ, 461, 933
  • Koenigl (1991) Koenigl, A. 1991, ApJ, 370, L39
  • Kruskal & Schwarzschild (1954) Kruskal, M. & Schwarzschild, M. 1954, Proceedings of the Royal Society of London Series A, 223, 348
  • Kulkarni & Romanova (2008) Kulkarni, A. K. & Romanova, M. M. 2008, MNRAS, 386, 673
  • Lesur et al. (2014) Lesur, G., Kunz, M. W., & Fromang, S. 2014, A&A, 566, A56
  • Lesur (2021) Lesur, G. R. J. 2021, A&A, 650, A35
  • Lovelace et al. (2010) Lovelace, R. V. E., Romanova, M. M., Ustyugova, G. V., & Koldoba, A. V. 2010, MNRAS, 408, 2083
  • Lynden-Bell & Pringle (1974) Lynden-Bell, D. & Pringle, J. E. 1974, MNRAS, 168, 603
  • Manara et al. (2023) Manara, C. F., Ansdell, M., Rosotti, G. P., et al. 2023, in Astronomical Society of the Pacific Conference Series, Vol. 534, Protostars and Planets VII, ed. S. Inutsuka, Y. Aikawa, T. Muto, K. Tomida, & M. Tamura, 539
  • Matt & Pudritz (2005) Matt, S. & Pudritz, R. E. 2005, MNRAS, 356, 167
  • McKee & Tan (2003) McKee, C. F. & Tan, J. C. 2003, ApJ, 585, 850
  • Meijerink & Spaans (2005) Meijerink, R. & Spaans, M. 2005, A&A, 436, 397
  • Newcomb (1961) Newcomb, W. A. 1961, Physics of Fluids, 4, 391
  • Offner & Chaban (2017) Offner, S. S. R. & Chaban, J. 2017, ApJ, 847, 104
  • Offner et al. (2009) Offner, S. S. R., Klein, R. I., McKee, C. F., & Krumholz, M. R. 2009, ApJ, 703, 131
  • Rea et al. (2024) Rea, D. G., Simon, J. B., Carrera, D., et al. 2024, arXiv e-prints, arXiv:2404.07265
  • Rebull et al. (2014) Rebull, L. M., Cody, A. M., Covey, K. R., et al. 2014, AJ, 148, 92
  • Romanova et al. (2018) Romanova, M. M., Blinova, A. A., Ustyugova, G. V., Koldoba, A. V., & Lovelace, R. V. E. 2018, New A, 62, 94
  • Romanova et al. (2021) Romanova, M. M., Koldoba, A. V., Ustyugova, G. V., et al. 2021, MNRAS, 506, 372
  • Romanova et al. (2011) Romanova, M. M., Ustyugova, G. V., Koldoba, A. V., & Lovelace, R. V. E. 2011, MNRAS, 416, 416
  • Rosen (2022) Rosen, A. L. 2022, ApJ, 941, 202
  • Rosen & Krumholz (2020) Rosen, A. L. & Krumholz, M. R. 2020, AJ, 160, 78
  • Rosen et al. (2016) Rosen, A. L., Krumholz, M. R., McKee, C. F., & Klein, R. I. 2016, MNRAS, 463, 2553
  • Rosen et al. (2012) Rosen, A. L., Krumholz, M. R., & Ramirez-Ruiz, E. 2012, ApJ, 748, 97
  • Seifried et al. (2017) Seifried, D., Walch, S., Girichidis, P., et al. 2017, MNRAS, 472, 4797
  • Shakura & Sunyaev (1973) Shakura, N. I. & Sunyaev, R. A. 1973, A&A, 24, 337
  • Sheehan et al. (2022) Sheehan, P. D., Tobin, J. J., Looney, L. W., & Megeath, S. T. 2022, ApJ, 929, 76
  • Shu (1977) Shu, F. H. 1977, ApJ, 214, 488
  • Siess et al. (2000) Siess, L., Dufour, E., & Forestini, M. 2000, A&A, 358, 593
  • Suin et al. (2024) Suin, P., Zavagno, A., Colman, T., et al. 2024, A&A, 682, A76
  • Takasao et al. (2018) Takasao, S., Tomida, K., Iwasaki, K., & Suzuki, T. K. 2018, ApJ, 857, 4
  • Takasao et al. (2022) Takasao, S., Tomida, K., Iwasaki, K., & Suzuki, T. K. 2022, ApJ, 941, 73
  • Thyng et al. (2016) Thyng, K. M., Greene, C. A., Hetland, R. D., Zimmerle, H. M., & DiMarco, S. F. 2016, Oceanography
  • Tobin et al. (2020) Tobin, J. J., Sheehan, P. D., Megeath, S. T., et al. 2020, ApJ, 890, 130
  • Tomida et al. (2013) Tomida, K., Tomisaka, K., Matsumoto, T., et al. 2013, ApJ, 763, 6
  • Ulrich (1976) Ulrich, R. K. 1976, ApJ, 210, 377
  • van der Velden (2020) van der Velden, E. 2020, The Journal of Open Source Software, 5, 2004
  • van der Velden et al. (2024) van der Velden, E., Robert, C., Batten, A., et al. 2024, 1313e/CMasher: v1.8.0
  • Van Rossum & Drake (2009) Van Rossum, G. & Drake, F. L. 2009, Python 3 Reference Manual (Scotts Valley, CA: CreateSpace)
  • Wagg & Broekgaarden (2024) Wagg, T. & Broekgaarden, F. 2024, The Software Citation Station
  • Wagg & Broekgaarden (2024) Wagg, T. & Broekgaarden, F. S. 2024, arXiv e-prints, arXiv:2406.04405
  • White & Hillenbrand (2004) White, R. J. & Hillenbrand, L. A. 2004, ApJ, 616, 998
  • Wolk et al. (2015) Wolk, S. J., Günther, H. M., Poppenhaeger, K., et al. 2015, AJ, 150, 145
  • Wolk et al. (2018) Wolk, S. J., Günther, H. M., Poppenhaeger, K., et al. 2018, AJ, 155, 99
  • Wurster et al. (2019) Wurster, J., Bate, M. R., & Price, D. J. 2019, MNRAS, 489, 1719
  • Yuan et al. (2012) Yuan, F., Bu, D., & Wu, M. 2012, ApJ, 761, 130
  • Zanni & Ferreira (2009) Zanni, C. & Ferreira, J. 2009, A&A, 508, 1117
  • Zanni & Ferreira (2013) Zanni, C. & Ferreira, J. 2013, A&A, 550, A99
  • Zhu et al. (2024) Zhu, Z., Stone, J. M., & Calvet, N. 2024, MNRAS, 528, 2883