Dust-Evacuated Zones Near Massive Stars: Consequences of Dust Dynamics on Star-forming Regions

Nadine H. Soliman TAPIR, Mailcode 350-17, California Institute of Technology, Pasadena, CA 91125, USA Philip F. Hopkins TAPIR, Mailcode 350-17, California Institute of Technology, Pasadena, CA 91125, USA Michael Y. Grudić Carnegie Observatories, 813 Santa Barbara St, Pasadena, CA 91101, USA
Abstract

Stars form within dense cores composed of both gas and dust within molecular clouds. However, despite the crucial role that dust plays in the star formation process, its dynamics is frequently overlooked, with the common assumption being a constant, spatially uniform dust-to-gas ratio and grain size spectrum. In this study, we introduce a set of radiation-dust-magnetohydrodynamic simulations of star forming molecular clouds from the STARFORGE project. These simulations expand upon the earlier radiation MHD models, which included cooling, individual star formation, and feedback. Notably, they explicitly address the dynamics of dust grains, considering radiation, drag, and Lorentz forces acting on a diverse size spectrum of live dust grains. We find that interactions between radiation and dust significantly influence the properties of gas surrounding and accreting onto massive stars. Specifically, we find that once stars exceed a certain mass threshold (2Msimilar-toabsent2subscript𝑀direct-product\sim 2M_{\odot}∼ 2 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT), their emitted radiation can evacuate dust grains from their vicinity, giving rise to a dust-suppressed zone of size 100similar-toabsent100\sim 100∼ 100 AU. Commencing during the early accretion stages and preceding the Main-sequence phase, this process results in a mass-dependent depletion in the accreted dust-to-gas (ADG) mass ratio within both the circumstellar disc and the star. We predict massive stars (10Mgreater-than-or-equivalent-toabsent10subscript𝑀direct-product\gtrsim 10M_{\odot}≳ 10 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT) would exhibit ADG ratios that are approximately one order of magnitude lower than that of their parent clouds. Consequently, stars, their discs, and circumstellar environments would display notable deviations in the abundances of elements commonly associated with dust grains, such as carbon and oxygen.

Star formation(1569) — Stellar abundances(1577) — Massive stars(732) — Interstellar dynamics(839) — Interstellar dust(836) — Interstellar dust extinction(837)

STARFORGE: Dust-Evacuated ZonesSTARFORGE: Dust-Evacuated Zones

1 Introduction

Star formation and stellar evolution are complex processes influenced by a multitude of physical mechanisms and environmental factors within Giant Molecular Clouds (GMCs) (McKee & Ostriker, 2007; Girichidis et al., 2020). The presence of magnetized, supersonic, and turbulent flows within these clouds drives strong density fluctuations, giving rise to regions with varying densities and spatial dimensions (Larson, 1981; Mac Low & Klessen, 2004). As these density fluctuations develop, some reach a critical point where their gravitational force begins to dominate, initiating protostellar collapse. The collapse of these gas and dust over-densities gives birth to protostars, which accrete nearby material and evolve into mature stars. During this process, they dynamically interact with the surrounding cloud through feedback mechanisms such as radiation, jets, radiatively-driven stellar winds, and supernova explosions (Krause et al., 2020).

Dust, an essential component of GMCs, heavily influences all of these processes. These particles, known as “grains”, are created as a byproduct of stellar evolution, within the atmospheres of evolved stars and in supernova remnants. Dust grains reprocess stellar radiation, absorbing far-ultraviolet (FUV), optical and infrared (IR) photons, re-emitting them in the IR and submillimeter wavelengths (Draine & Lee, 1984; Mathis, 1990; Li & Draine, 2001; Tielens, 2005). Furthermore, they scatter background radiation and emit a thermal continuum in the IR. In addition to its radiative interactions, dust governs the thermodynamics and chemistry of gas within the ISM, exerting a substantial influence on the intricate processes triggering star formation (Whittet et al., 1993; Salpeter, 1977; Weingartner & Draine, 2001, 2001; Draine, 2003; Dorschner, 2003; Spitzer Jr, 2008; Watanabe & Kouchi, 2008; Minissale et al., 2016). Moreover, dust acts as a reservoir, confining heavy elements within a solid phase, which can subsequently become integrated into stars and planets, thereby significantly influencing their overall compositions. Once luminous sources form within the GMC, the presence of dust can also initiate radiatively driven outflows in the surrounding regions (Murray et al., 2005; King & Pounds, 2015; Höfner & Olofsson, 2018).

In addition, the dynamic interplay between dust and gas is pivotal in sha** the evolution of star-forming regions. Turbulence within the cloud can generate variations in the dust density on scales comparable to the turbulent eddy turnover scale, leading to substantial fluctuations in the dust-to-gas (DTG) ratio within prestellar cores (Thoraval et al., 1997, 1999; Abergel et al., 2002; Flagey et al., 2009; Boogert et al., 2013). As demonstrated by Hopkins (2014), this phenomenon has significant implications within large GMCs, where dust dynamics alone could potentially lead to stellar populations that exhibit notable variations in abundances of elements commonly found in dust grains, including CNO, Si, Mg and Fe (Hopkins & Conroy, 2017).

Despite its well-recognized importance, most simulations that model star-formation often assume a perfect coupling between the dust and the surrounding gas. This treatment involves considering both components as moving together, with dust essentially acting as an “additional opacity" to the gas. However, the dynamics of dust is far more complex. Dust particles, ranging from a few angstroms to several micrometres in size, are inherently aerodynamic and often charged. When dust grains are accelerated by forces like radiation pressure, they drift through the gas, and the imparted momentum couples to the surrounding gas through electromagnetic and drag forces, redistributing momentum in the environment. The efficiency of this coupling relies on both the characteristics of the grains and the surrounding environment. For typical densities of GMCs, where nH102cm3similar-tosubscript𝑛Hsuperscript102superscriptcm3n_{\rm H}\sim 10^{2}\,\rm cm^{-3}italic_n start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT ∼ 10 start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT, the collisional interactions between dust and gas are relatively infrequent, resulting in a relatively weak coupling of the dust dynamics with that of the gas. This decoupling between grain and gas populations has notably been observed, particularly in the case of larger dust grains (Krüger et al., 2001; Frisch & Slavin, 2003; Meisel et al., 2002; Altobelli et al., 2006, 2007; Poppe et al., 2010). To improve our understanding of dust dynamics within this environment, this study serves as an initial investigation that relaxes some assumptions regarding the perfect coupling of gas and dust dynamics.

In this work, we present the first radiation-dust-magnetohydrodynamic (RDMHD) simulations of star-forming molecular clouds that explicitly account for dust dynamics. We utilize simulations from the STAR FORmation in Gaseous Environments (STARFORGE) project, which provide a comprehensive representation of individual star formation. This encompasses the stages of proto-stellar collapse, subsequent accretion and stellar feedback, main-sequence evolution, and stellar dynamics, with a thorough consideration of the relevant physical processes (as described in Grudić et al. (2022)). Within the context of this paper, our focus is directed towards investigating the impacts of explicit dust-gas radiation dynamics on the properties of the stellar populations that emerge within the cloud.

The paper is structured as follows: In Section 2, a concise outline of the code is presented (for a detailed description of the numerical methods and implementations, refer to Grudić et al. (2021)) along with a description of the initial conditions (ICs) of the simulations. We discuss the results obtained from our fiducial simulation runs in Section 3 and compare them to runs with simplified dust physics. Finally, we discuss the implications of our findings in Section 4 and present our conclusions in Section 5.

2 Simulations

2.1 STARFORGE Physics

We run a set of RDMHD STARFORGE simulations to study star formation in GMCs, using the GIZMO code (Hopkins, 2015). These simulations adopt the complete physics setup from the “full” STARFORGE model, as described in Grudić et al. (2022). For solving the equations of ideal magnetohydrodynamics (MHD), we rely on the GIZMO Meshless Finite Mass MHD solver, detailed in Hopkins & Raives (2016) and Hopkins (2016). In this study, we extend the standard STARFORGE physics by explicitly modeling the dynamics of the dust particles, where the details of the implementation are outlined in Section 2.2.

We utilize the GIZMO meshless frequency-integrated M1 solver (Lupi et al., 2017, 2018; Hopkins & Grudić, 2019a; Hopkins et al., 2020; Grudić et al., 2021; Hopkins & Grudić, 2019a; Hopkins et al., 2020) to evolve the time-dependent, frequency-integrated radiative transfer (RT) equation adopting a reduced speed of light c=30kms1𝑐30superscriptkms1c=30\,\rm kms^{-1}italic_c = 30 roman_kms start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT. The radiation is discretised into five distinct frequency bands, which cover a range extending from the Lyman continuum to the far infrared (FIR). As radiation interacts with matter, it triggers processes such as photoionisation, photodissociation, photoelectric heating, and dust absorption. Additionally, the matter undergoes radiative cooling and heating, accounting for metal lines, molecular lines, fine structure lines, as well as continuum and dust collisional processes, as outlined in Hopkins et al. (2023). An especially noteworthy feature of our radiative transfer approach is the direct coupling of each radiation band to the dust particles. A more comprehensive description of this integration is presented in Section 2.2. The radiation field is initialised with an external heating source at the boundaries, representing the interstellar radiation field (ISRF). As local sources (stars) emerge, they also contribute to the radiation field.

Individual stars in the simulations are represented by sink particles. These sink particles arise from gas cells that meet the criteria for runaway gravitational collapse and follow the protostellar evolution model introduced by McKee & Offner (2010). Sink particles can then accrete bound gas and dust elements, where dust accretion adheres to the same critera as gas accretion. Each sink particle separately tracks the quantities of both dust and gas it has accumulated since its initial sink formation. To ensure realistic accretion irrespective of resolution, the sink particles first accrete particles onto an unresolved disk reservoir, the material in which is then smoothly accreted onto the sink particle.

The sink particles undergo growth through accretion while progressing along the main sequence. Their luminosity and radius follow the relationships outlined in Tout et al. (1996), and they emit a black-body spectrum with an effective temperature of Teff=5780,K(L/R2)1/4subscript𝑇eff5780KsuperscriptsubscriptLsuperscriptsubscriptR214T_{\rm eff}=5780,\rm K\left(L_{\star}/R_{\star}^{2}\right)^{1/4}italic_T start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT = 5780 , roman_K ( roman_L start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT / roman_R start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 1 / 4 end_POSTSUPERSCRIPT. Beyond radiation, sinks interact with their surrounding medium via protostellar jets, stellar winds, and the potential occurrence of supernovae. To calculate gravitational forces, we employ an adaptive gravitational softening approach, which spans a range extending down to approximately 2×105similar-toabsent2superscript105\sim 2\times 10^{-5}∼ 2 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT AU for the gas cells. Additionally, a fixed softening of 6666 AU is utilized for the dust and sink particles. For a more comprehensive description of the modelled physics and numerical techniques, readers are encouraged to refer to Grudić et al. (2021, 2022).

2.2 Dust Physics

A detailed description of the numerical methods used for dust modelling in our simulations, albeit focussing on more idealised scenarios, can be found in Hopkins & Lee (2016); Lee et al. (2017); Moseley et al. (2019); Soliman & Hopkins (2023). Following a Monte Carlo sampling approach, we depict dust grains in our simulation as "super-particles" (Carballido et al., 2008; Johansen et al., 2009; Bai & Stone, 2010; Pan et al., 2011; McKinnon et al., 2018). Each simulated “dust particle” encapsulates an ensemble of dust grains characterized by similar attributes such as grain size (ϵgrain)subscriptitalic-ϵgrain(\epsilon_{\text{grain}})( italic_ϵ start_POSTSUBSCRIPT grain end_POSTSUBSCRIPT ), mass (mgrain)subscript𝑚grain(m_{\text{grain}})( italic_m start_POSTSUBSCRIPT grain end_POSTSUBSCRIPT ), and charge (qgrain)subscript𝑞grain(q_{\text{grain}})( italic_q start_POSTSUBSCRIPT grain end_POSTSUBSCRIPT ) determined through live collisional, photoelectric, and cosmic ray charging (Draine & Sutin, 1987; Tielens, 2005).

The motion of each dust grain is governed by the following equation:

d𝐯ddtdsubscript𝐯𝑑dt\displaystyle\frac{\mathrm{d}{\bf v}_{d}}{\mathrm{dt}}divide start_ARG roman_d bold_v start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG roman_dt end_ARG =𝐚gas,dust+𝐚grav+𝐚radabsentsubscript𝐚gasdustsubscript𝐚gravsubscript𝐚rad\displaystyle={\bf a}_{\rm gas,\,dust}+{\bf a}_{\rm grav}+{\bf a}_{\rm rad}= bold_a start_POSTSUBSCRIPT roman_gas , roman_dust end_POSTSUBSCRIPT + bold_a start_POSTSUBSCRIPT roman_grav end_POSTSUBSCRIPT + bold_a start_POSTSUBSCRIPT roman_rad end_POSTSUBSCRIPT (1)
=𝐰sts𝐰s×𝐁^tL+𝐠+πϵgrain2mgraincQext𝐆rad,absentsubscript𝐰𝑠subscript𝑡𝑠subscript𝐰𝑠^𝐁subscript𝑡𝐿𝐠𝜋superscriptsubscriptitalic-ϵgrain2subscript𝑚grain𝑐subscriptdelimited-⟨⟩𝑄extsubscript𝐆rad\displaystyle=-\frac{{\bf w}_{s}}{t_{s}}-\frac{{\bf w}_{s}\times\hat{\bf B}}{t% _{L}}+{\bf g}+\frac{\pi\,\epsilon_{\rm grain}^{2}}{m_{\rm grain}\,c}\,\langle Q% \rangle_{\rm ext}\,{\bf G}_{\rm rad},= - divide start_ARG bold_w start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_ARG start_ARG italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_ARG - divide start_ARG bold_w start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT × over^ start_ARG bold_B end_ARG end_ARG start_ARG italic_t start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT end_ARG + bold_g + divide start_ARG italic_π italic_ϵ start_POSTSUBSCRIPT roman_grain end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_m start_POSTSUBSCRIPT roman_grain end_POSTSUBSCRIPT italic_c end_ARG ⟨ italic_Q ⟩ start_POSTSUBSCRIPT roman_ext end_POSTSUBSCRIPT bold_G start_POSTSUBSCRIPT roman_rad end_POSTSUBSCRIPT ,

where 𝐯dsubscript𝐯𝑑{\bf v}_{d}bold_v start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT represents the velocity of the grain; 𝐚gas,dust=𝐰s/ts𝐰s×𝐁^/tLsubscript𝐚gasdustsubscript𝐰𝑠subscript𝑡𝑠subscript𝐰𝑠^𝐁subscript𝑡𝐿{\bf a}_{\rm gas,\,dust}=-{{\bf w}_{s}}/{t_{s}}-{{\bf w}_{s}\times\hat{\bf B}}% /{t_{L}}bold_a start_POSTSUBSCRIPT roman_gas , roman_dust end_POSTSUBSCRIPT = - bold_w start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT / italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - bold_w start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT × over^ start_ARG bold_B end_ARG / italic_t start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT takes into account the forces exerted by the gas on the dust, including drag (quantified by the "stop** time" tssubscript𝑡𝑠t_{s}italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT) and Lorentz forces (characterised by the gyro/Larmor time tLmgrainc/|qgrain,𝐁|t_{L}\equiv m_{\rm grain}c/|q_{\rm grain},{\bf B}|italic_t start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ≡ italic_m start_POSTSUBSCRIPT roman_grain end_POSTSUBSCRIPT italic_c / | italic_q start_POSTSUBSCRIPT roman_grain end_POSTSUBSCRIPT , bold_B |); 𝐰s𝐯d𝐮gsubscript𝐰𝑠subscript𝐯𝑑subscript𝐮𝑔{\bf w}_{s}\equiv{\bf v}_{d}-{\bf u}_{g}bold_w start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ≡ bold_v start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT - bold_u start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT corresponds to the drift velocity of a dust grain relative to the gas velocity 𝐮gsubscript𝐮𝑔{\bf u}_{g}bold_u start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT at the same position 𝐱𝐱{\bf x}bold_x; 𝐁𝐁{\bf B}bold_B denotes the local magnetic field; 𝐚grav=𝐠subscript𝐚grav𝐠{\bf a}_{\rm grav}={\bf g}bold_a start_POSTSUBSCRIPT roman_grav end_POSTSUBSCRIPT = bold_g is the gravitational force due to a local gravitational field; and 𝐚radsubscript𝐚rad{\bf a}_{\rm rad}bold_a start_POSTSUBSCRIPT roman_rad end_POSTSUBSCRIPT is the force due to an incident radiation field 𝐆rad𝐅rad𝐯d(erad𝕀+rad)subscript𝐆radsubscript𝐅radsubscript𝐯𝑑subscript𝑒rad𝕀subscriptrad{\bf G}_{\rm rad}\equiv{\bf F}_{\rm rad}-{\bf v}_{d}\cdot(e_{\rm rad}\,\mathbb% {I}+\mathbb{P}_{\rm rad})bold_G start_POSTSUBSCRIPT roman_rad end_POSTSUBSCRIPT ≡ bold_F start_POSTSUBSCRIPT roman_rad end_POSTSUBSCRIPT - bold_v start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ⋅ ( italic_e start_POSTSUBSCRIPT roman_rad end_POSTSUBSCRIPT blackboard_I + blackboard_P start_POSTSUBSCRIPT roman_rad end_POSTSUBSCRIPT ) in terms of the radiation flux/energy density/pressure density 𝐅radsubscript𝐅rad{\bf F}_{\rm rad}bold_F start_POSTSUBSCRIPT roman_rad end_POSTSUBSCRIPT, eradsubscript𝑒rade_{\rm rad}italic_e start_POSTSUBSCRIPT roman_rad end_POSTSUBSCRIPT, radsubscriptrad\mathbb{P}_{\rm rad}blackboard_P start_POSTSUBSCRIPT roman_rad end_POSTSUBSCRIPT for a grain of size ϵgrainsubscriptitalic-ϵgrain\epsilon_{\rm grain}italic_ϵ start_POSTSUBSCRIPT roman_grain end_POSTSUBSCRIPT and mass mgrain(4π/3)ρ¯grainiϵgrain3subscript𝑚grain4𝜋3superscriptsubscript¯𝜌grain𝑖superscriptsubscriptitalic-ϵgrain3m_{\rm grain}\equiv(4\pi/3)\,\bar{\rho}_{\rm grain}^{\,i}\,\epsilon_{\rm grain% }^{3}italic_m start_POSTSUBSCRIPT roman_grain end_POSTSUBSCRIPT ≡ ( 4 italic_π / 3 ) over¯ start_ARG italic_ρ end_ARG start_POSTSUBSCRIPT roman_grain end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT italic_ϵ start_POSTSUBSCRIPT roman_grain end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT, where ρ¯grainisuperscriptsubscript¯𝜌grain𝑖\bar{\rho}_{\rm grain}^{\,i}over¯ start_ARG italic_ρ end_ARG start_POSTSUBSCRIPT roman_grain end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT is the internal grain mass density, and dimensionless absorption+scattering efficiency Qextsubscriptdelimited-⟨⟩𝑄ext\langle Q\rangle_{\rm ext}⟨ italic_Q ⟩ start_POSTSUBSCRIPT roman_ext end_POSTSUBSCRIPT; c𝑐citalic_c is the speed of light.

Our radiation treatment closely follows the methodology described in Hopkins et al. (2022); Soliman & Hopkins (2023), which employed a gray-band opacity assumption. In our current study, we build upon this approach by introducing a five-band opacity treatment for each radiation band evolved by the RT solver. We determine an effective opacity, expressed as Qext(ϵgrain,λeff)=min(2πϵgrain/λeff,1)subscriptdelimited-⟨⟩𝑄extsubscriptitalic-ϵgrainsubscript𝜆eff2𝜋subscriptitalic-ϵgrainsubscript𝜆eff1\langle Q\rangle_{\rm ext}(\epsilon_{\rm grain},\lambda_{\rm eff})=\min\left(2% \pi\epsilon_{\rm grain}/\lambda_{\rm eff},1\right)⟨ italic_Q ⟩ start_POSTSUBSCRIPT roman_ext end_POSTSUBSCRIPT ( italic_ϵ start_POSTSUBSCRIPT roman_grain end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT ) = roman_min ( 2 italic_π italic_ϵ start_POSTSUBSCRIPT roman_grain end_POSTSUBSCRIPT / italic_λ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT , 1 ), where λeffλminλmaxsubscript𝜆effsubscript𝜆subscript𝜆\lambda_{\rm eff}\equiv\sqrt{\lambda_{\min}\lambda_{\max}}italic_λ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT ≡ square-root start_ARG italic_λ start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT end_ARG, representing the geometric mean of the wavelength boundaries for the respective radiation bands.

The radiation pressure force on the dust is determined through the M1 radiative transfer method described above, encompassing terms up to 𝒪(v2/c2)𝒪superscript𝑣2superscript𝑐2\mathcal{O}(v^{2}/c^{2})caligraphic_O ( italic_v start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_c start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ): terad+𝐅rad=Rdust𝐯d𝐆rad/c2subscript𝑡subscript𝑒radsubscript𝐅radsubscript𝑅dustsubscript𝐯𝑑subscript𝐆radsuperscript𝑐2\partial_{t}e_{\rm rad}+\nabla\cdot{\bf F}_{\rm rad}=-R_{\rm dust}\,{\bf v}_{d% }\cdot{\bf G}_{\rm rad}/c^{2}∂ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT roman_rad end_POSTSUBSCRIPT + ∇ ⋅ bold_F start_POSTSUBSCRIPT roman_rad end_POSTSUBSCRIPT = - italic_R start_POSTSUBSCRIPT roman_dust end_POSTSUBSCRIPT bold_v start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ⋅ bold_G start_POSTSUBSCRIPT roman_rad end_POSTSUBSCRIPT / italic_c start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, t𝐅rad+c2rad=Rdust𝐆radsubscript𝑡subscript𝐅radsuperscript𝑐2subscriptradsubscript𝑅dustsubscript𝐆rad\partial_{t}{\bf F}_{\rm rad}+c^{2}\,\nabla\cdot\mathbb{P}_{\rm rad}=-R_{\rm dust% }\,{\bf G}_{\rm rad}∂ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT bold_F start_POSTSUBSCRIPT roman_rad end_POSTSUBSCRIPT + italic_c start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ∇ ⋅ blackboard_P start_POSTSUBSCRIPT roman_rad end_POSTSUBSCRIPT = - italic_R start_POSTSUBSCRIPT roman_dust end_POSTSUBSCRIPT bold_G start_POSTSUBSCRIPT roman_rad end_POSTSUBSCRIPT. Specifically, it involves equations that govern the temporal evolution of quantities related to the radiation field, where the absorption and scattering coefficients, denoted as Rdustsubscript𝑅dustR_{\rm dust}italic_R start_POSTSUBSCRIPT roman_dust end_POSTSUBSCRIPT, are calculated locally from the dust grain distribution.

However, resolving the photon mean free path in global simulations is not always feasible (Hopkins & Grudić, 2019b; Krumholz, 2018). To address this limitation, we apply a correction factor as outlined by Grudić et al. (2021) to estimate radiation flux. Additionally, in this work we account for multiple scattering, where photons absorbed by dust are presumed to be re-emitted within the infrared band.

To examine the effects of this implementation, we conduct a series of numerical experiments, systematically introducing and removing the correction factors for each radiation band. Our analysis reveals that while these adjustments can influence the results quantitatively, their effects generally remain within the interquartile range observed in this study.

For all forces originating from the interaction between gas and dust, denoted agas,dustsubscript𝑎gasdusta_{\rm gas,dust}italic_a start_POSTSUBSCRIPT roman_gas , roman_dust end_POSTSUBSCRIPT, an equal and opposite force acts on the gas (referred to as “back-reaction”). For the physical conditions under consideration, the drag experienced by the dust is modeled using the Epstein drag formulation, expressed as:

tssubscript𝑡𝑠\displaystyle t_{s}italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT πγ8ρ¯grainiϵgrainρgcs(1+9πγ128|𝐰s|2cs2)1/2,absent𝜋𝛾8superscriptsubscript¯𝜌grain𝑖subscriptitalic-ϵgrainsubscript𝜌𝑔subscript𝑐𝑠superscript19𝜋𝛾128superscriptsubscript𝐰𝑠2superscriptsubscript𝑐𝑠212\displaystyle\equiv\sqrt{\frac{\pi\gamma}{8}}\frac{\bar{\rho}_{\rm grain}^{\,i% }\,\epsilon_{\rm grain}}{\rho_{g}\,c_{s}}\,\bigg{(}1+\frac{9\pi\gamma}{128}% \frac{|{\bf w}_{s}|^{2}}{c_{s}^{2}}\bigg{)}^{-1/2},≡ square-root start_ARG divide start_ARG italic_π italic_γ end_ARG start_ARG 8 end_ARG end_ARG divide start_ARG over¯ start_ARG italic_ρ end_ARG start_POSTSUBSCRIPT roman_grain end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT italic_ϵ start_POSTSUBSCRIPT roman_grain end_POSTSUBSCRIPT end_ARG start_ARG italic_ρ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_ARG ( 1 + divide start_ARG 9 italic_π italic_γ end_ARG start_ARG 128 end_ARG divide start_ARG | bold_w start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) start_POSTSUPERSCRIPT - 1 / 2 end_POSTSUPERSCRIPT , (2)

where γ𝛾\gammaitalic_γ is the adiabatic index, ρgsubscript𝜌𝑔\rho_{g}italic_ρ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT is the gas density, and cssubscript𝑐𝑠c_{s}italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT is the local sound speed. Additionally, Coulomb drag is considered, though it typically constitutes a minor correction.

Furthermore, the thermochemistry of the medium is influenced by the presence of dust through both indirect effects, arising from extinction by grains, and direct effects, through dust heating and cooling terms (refer to Grudić et al. (2022) for comprehensive details). These effects depend on the DTG ratio and/or mean grain size. Instead of assuming fixed values, we estimate these parameters by interpolating the local distribution of dust grains to the gas cells, ensuring a self-consistent consideration of the thermodynamics involved. We examine the implications of our explicit dust treatment on the thermodynamics of star-forming regions and delve into the specific influence of varying dust properties, such as grain size, in more detail in Soliman & Hopkins (in prep.).

The simulations at hand do not resolve the detailed physics of the generation and collimation of protostellar jets. Instead, jets are introduced by spawning high velocity gas cells within a narrow cone (see Grudić et al. (2022) for implementation details). The cone’s radius encompasses a few gas cells within the sink radius, and is thus poorly resolved. To mitigate potential spurious effects, such as the ejection of dust grains by the spawned jet particles, we temporarily disable the interaction between dust particles and newly spawned cells.This involves interpolating the gas properties from non-jet cells only to dust for a specific duration, allowing the jets to escape the poorly resolved sink radius. We experimented with the duration of the decoupling, and find that results converge as long as the decoupling persists until the jets can escape the sink radius or become resolved. We emphasize that we adopt this approach as a precautionary measure to avoid overestimating the extent of dust evacuation. Allowing the interaction to proceed would only lead to further dust ejection near sink particles. Consequently, this approach does not qualitatively affect our findings; instead, it strengthens our conclusions.

2.3 Initial conditions

For our fiducial simulations, we consider an initially uniform-density turbulent molecular cloud with a mass of Mcloud=2×103Msubscript𝑀cloud2superscript103subscriptMdirect-productM_{\rm cloud}=2\times 10^{3}\rm M_{\odot}italic_M start_POSTSUBSCRIPT roman_cloud end_POSTSUBSCRIPT = 2 × 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT and a radius of R=3pc𝑅3pcR=3\rm pcitalic_R = 3 roman_p roman_c. The cloud is enveloped within a 30pc30pc30\rm pc30 roman_p roman_c periodic box with an ambient medium with density 103similar-toabsentsuperscript103\sim 10^{3}∼ 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT times lower than that of the cloud. The initial velocity distribution follows Gaussian random field with with an initial virial parameter αturb=5σ2R/(3GMcloud)=2subscript𝛼turb5superscript𝜎2𝑅3𝐺subscript𝑀cloud2\alpha_{\rm turb}=5\sigma^{2}R/(3GM_{\rm cloud})=2italic_α start_POSTSUBSCRIPT roman_turb end_POSTSUBSCRIPT = 5 italic_σ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_R / ( 3 italic_G italic_M start_POSTSUBSCRIPT roman_cloud end_POSTSUBSCRIPT ) = 2. The initial magnetic field configuration is uniform, and set to establish a mass-to-flux ratio equivalent to 4.2 times the critical value within the cloud. For typical gas cells, the mass resolution in our simulations is Δm103Msimilar-toΔ𝑚superscript103subscriptMdirect-product\Delta m\sim 10^{-3}\rm M_{\odot}roman_Δ italic_m ∼ 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, while for the grain super-particles, we refine it to 2.5×106M2.5superscript106subscriptMdirect-product2.5\times 10^{-6}\rm M_{\odot}2.5 × 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT (4×4\times4 × higher resolution for the dust). Furthermore, cells associated with protostellar jets and stellar winds have a higher mass resolution of 104Msuperscript104subscriptMdirect-product10^{-4}\rm M_{\odot}10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT. In addition, we extend our sample to include a larger cloud configuration with M=2×104M𝑀2superscript104subscriptMdirect-productM=2\times 10^{4}\rm M_{\odot}italic_M = 2 × 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT and radius R=10pc𝑅10pcR=10\rm pcitalic_R = 10 roman_p roman_c also enveloped within a 10×\times× larger box filled with diffuse material. For the larger cloud, we use a coarser mass resolution, scaled down by a factor of 10.

The dust component is initialized with a net zero drift velocity relative to the surrounding gas. It is spatially distributed according to an initially uniform DTG ratio ρd0=μdgρg0superscriptsubscript𝜌𝑑0superscript𝜇dgsuperscriptsubscript𝜌𝑔0\rho_{d}^{0}=\mu^{\rm dg}\rho_{g}^{0}italic_ρ start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT = italic_μ start_POSTSUPERSCRIPT roman_dg end_POSTSUPERSCRIPT italic_ρ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT, where μdg=0.01superscript𝜇dg0.01\mu^{\rm dg}=0.01italic_μ start_POSTSUPERSCRIPT roman_dg end_POSTSUPERSCRIPT = 0.01 reflects the galactic value. The grain sizes follow an empirical Mathis, Rumpl, & Nordsieck (MRN) power-law distribution with a differential number density dNd/dϵgrainϵgrain3.5proportional-to𝑑subscript𝑁d𝑑subscriptitalic-ϵgrainsuperscriptsubscriptitalic-ϵgrain3.5dN_{\rm d}/d\epsilon_{\rm grain}\propto\epsilon_{\rm grain}^{-3.5}italic_d italic_N start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT / italic_d italic_ϵ start_POSTSUBSCRIPT roman_grain end_POSTSUBSCRIPT ∝ italic_ϵ start_POSTSUBSCRIPT roman_grain end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 3.5 end_POSTSUPERSCRIPT (Mathis et al., 1977). The distribution spans a dynamic range of ϵgrainmax=100ϵgrainminsuperscriptsubscriptitalic-ϵgrainmax100superscriptsubscriptitalic-ϵgrainmin\epsilon_{\rm grain}^{\rm max}=100\epsilon_{\rm grain}^{\rm min}italic_ϵ start_POSTSUBSCRIPT roman_grain end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_max end_POSTSUPERSCRIPT = 100 italic_ϵ start_POSTSUBSCRIPT roman_grain end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_min end_POSTSUPERSCRIPT, with ϵgrainmax=0.1μmsuperscriptsubscriptitalic-ϵgrainmax0.1𝜇m\epsilon_{\rm grain}^{\rm max}=0.1\rm\mu mitalic_ϵ start_POSTSUBSCRIPT roman_grain end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_max end_POSTSUPERSCRIPT = 0.1 italic_μ roman_m for our fiducial simulations. We adopt the classic MRN mixture of carbonaceous and silicate composition, assuming a uniform internal density and composition across different grain sizes. This corresponds to grains with a sublimation temperature of approximately 1500 K and an internal density of ρ¯graini2.25g/cm3similar-tosuperscriptsubscript¯𝜌grain𝑖2.25gsuperscriptcm3\bar{\rho}_{\rm grain}^{\,i}\sim 2.25\,\rm g/cm^{3}over¯ start_ARG italic_ρ end_ARG start_POSTSUBSCRIPT roman_grain end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT ∼ 2.25 roman_g / roman_cm start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT. We do not model grain growth or destruction; consequently, we adhere to a fixed size distribution, maintaining constant sizes for particles throughout the simulation.

We provide a summary of the initial conditions for all simulations discussed in this work in Table LABEL:table.

Name Mcloudsubscript𝑀cloudM_{\rm cloud}italic_M start_POSTSUBSCRIPT roman_cloud end_POSTSUBSCRIPT [MsubscriptMdirect-product\rm M_{\odot}roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT] Rcloudsubscript𝑅cloudR_{\rm cloud}italic_R start_POSTSUBSCRIPT roman_cloud end_POSTSUBSCRIPT [pc] ϵgrainmaxsuperscriptsubscriptitalic-ϵgrainmax\epsilon_{\rm grain}^{\rm max}italic_ϵ start_POSTSUBSCRIPT roman_grain end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_max end_POSTSUPERSCRIPT [μm𝜇m\rm\mu mitalic_μ roman_m] ΔmΔ𝑚\Delta mroman_Δ italic_m [MsubscriptMdirect-product\rm M_{\odot}roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT] Notes
m2e3_0.1 2×1032superscript1032\times 10^{3}2 × 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT 3 0.1 102superscript10210^{-2}10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT fiducial run
m2e3_0.1_hires 2×1032superscript1032\times 10^{3}2 × 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT 3 0.1 103superscript10310^{-3}10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT ×\times× 10 finer resolution
m2e3_0.1_nomag 2×1032superscript1032\times 10^{3}2 × 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT 3 0.1 102superscript10210^{-2}10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT no Lorentz Forces on grains
m2e3_0.1_norad 2×1032superscript1032\times 10^{3}2 × 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT 3 0.1 102superscript10210^{-2}10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT no radiation pressure forces on grains
m2e3_0.1_pass 2×1032superscript1032\times 10^{3}2 × 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT 3 0.1 102superscript10210^{-2}10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT grains only feel drag & do not exert a back-reaction force
m2e3_0.1_nofb 2×1032superscript1032\times 10^{3}2 × 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT 3 0.1 102superscript10210^{-2}10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT no stellar winds or protostellar jets
m2e3_1 2×1032superscript1032\times 10^{3}2 × 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT 3 1.0 102superscript10210^{-2}10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT ×\times× 10 larger grains
m2e3_10 2×1032superscript1032\times 10^{3}2 × 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT 3 10 102superscript10210^{-2}10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT ×100absent100\times 100× 100 larger grains
m2e3_1_hires 2×1032superscript1032\times 10^{3}2 × 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT 3 1.0 103superscript10310^{-3}10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT ×\times× 10 larger grains & ×\times×10 finer resolution
m2e3_10_hires 2×1032superscript1032\times 10^{3}2 × 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT 3 10 103superscript10310^{-3}10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT ×100absent100\times 100× 100 larger grains & ×\times×10 coarser resolution
m2e4_0.1 2×1042superscript1042\times 10^{4}2 × 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT 10 0.1 102superscript10210^{-2}10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT fiducial run
m2e4_0.1_hires 2×1042superscript1042\times 10^{4}2 × 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT 10 0.1 103superscript10310^{-3}10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT ×10absent10\times 10× 10 finer resolution
m2e4_0.1_nomag 2×1042superscript1042\times 10^{4}2 × 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT 10 0.1 102superscript10210^{-2}10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT no Lorentz forces on grains
m2e4_0.1_norad 2×1042superscript1042\times 10^{4}2 × 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT 10 0.1 102superscript10210^{-2}10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT no radiation pressure forces on grains
m2e4_0.1_pass 2×1042superscript1042\times 10^{4}2 × 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT 10 0.1 102superscript10210^{-2}10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT grains only feel drag & do not exert a back-reaction force
m2e4_1 2×1042superscript1042\times 10^{4}2 × 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT 10 1.0 102superscript10210^{-2}10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT ×\times×10 larger grains
m2e4_10 2×1042superscript1042\times 10^{4}2 × 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT 10 10 102superscript10210^{-2}10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT ×\times×100 larger grains
Table 1: The initial conditions for the simulations used in this study. The columns include: (1) Simulation name. (2) Cloud mass Mcloudsubscript𝑀cloudM_{\rm cloud}italic_M start_POSTSUBSCRIPT roman_cloud end_POSTSUBSCRIPT. (3) Cloud Radius Rcloudsubscript𝑅cloudR_{\rm cloud}italic_R start_POSTSUBSCRIPT roman_cloud end_POSTSUBSCRIPT. (4) Maximum grain size ϵgrainmaxsuperscriptsubscriptitalic-ϵgrainmax\epsilon_{\rm grain}^{\rm max}italic_ϵ start_POSTSUBSCRIPT roman_grain end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_max end_POSTSUPERSCRIPT. (5) Mass resolution ΔmΔ𝑚\Delta mroman_Δ italic_m. (6) Notes indicating the main variations from the fiducial run.

3 Results

3.1 Cloud Morphology

In Figure 1, we present the 2D integrated surface density of gas, denoted as ΣgassubscriptΣgas\Sigma_{\rm gas}roman_Σ start_POSTSUBSCRIPT roman_gas end_POSTSUBSCRIPT displayed on the left and dust, denoted as ΣdustsubscriptΣdust\Sigma_{\rm dust}roman_Σ start_POSTSUBSCRIPT roman_dust end_POSTSUBSCRIPT shown on the right. These visualizations are from our m2e3_0.1_hires simulation, which employs our comprehensive physics model to simulate a cloud with an initial mass of Mcloud2×103Msimilar-tosubscript𝑀cloud2superscript103subscriptMdirect-productM_{\rm cloud}\sim 2\times 10^{3}\rm M_{\odot}italic_M start_POSTSUBSCRIPT roman_cloud end_POSTSUBSCRIPT ∼ 2 × 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT evolved for a duration of t3.5similar-to𝑡3.5t\sim 3.5italic_t ∼ 3.5 Myrs. We note that the dust surface density has been scaled by a factor of 1/μdg1superscript𝜇dg1/\mu^{\rm dg}1 / italic_μ start_POSTSUPERSCRIPT roman_dg end_POSTSUPERSCRIPT to facilitate a straightforward comparison between the distributions of gas and dust. Sink particles, representing stars in the system, are portrayed as circles with their sizes proportional to their respective masses. The initially spherically uniform cloud undergoes a process of gravitational collapse and fragmentation. This leads to the emergence of a stellar cluster near the central region with a small number of sinks dispersed at greater distances from the cluster’s center. When we compare the dust and gas distributions on parsec scales, we observe that the distribution of the dust aligns with that of the gas. This indicates that, on large scales, the dynamics of dust and gas are effectively well-coupled.

Refer to caption
Figure 1: The 2D integrated surface density of gas ΣgassubscriptΣgas\Sigma_{\rm gas}roman_Σ start_POSTSUBSCRIPT roman_gas end_POSTSUBSCRIPT (left) and dust ΣdustΣdust\Sigma-{\rm dust}roman_Σ - roman_dust (right) at t3.5similar-to𝑡3.5t\sim 3.5italic_t ∼ 3.5 Myrs for the m2e3_0.1_hires simulation, corresponding to a cloud with Mcloud2×103Msimilar-tosubscript𝑀cloud2superscript103subscriptMdirect-productM_{\rm cloud}\sim 2\times 10^{3}\rm M_{\rm\odot}italic_M start_POSTSUBSCRIPT roman_cloud end_POSTSUBSCRIPT ∼ 2 × 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, Δm103Msimilar-toΔ𝑚superscript103subscriptMdirect-product\Delta m\sim 10^{-3}\rm M_{\rm\odot}roman_Δ italic_m ∼ 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT resolution, and ϵgrainmax=0.1μmsuperscriptsubscriptitalic-ϵgrainmax0.1𝜇m\epsilon_{\rm grain}^{\rm max}=0.1\rm\mu mitalic_ϵ start_POSTSUBSCRIPT roman_grain end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_max end_POSTSUPERSCRIPT = 0.1 italic_μ roman_m. This run employs our full physics package and accounts for explicit dust dynamics, including physical phenomena such as dust drag, dust back-reaction, Lorentz forces acting on dust grains, and the explicit computation of dust opacity from grain particles. Note that the dust surface density is scaled by 1/μdg1superscript𝜇dg1/\mu^{\rm dg}1 / italic_μ start_POSTSUPERSCRIPT roman_dg end_POSTSUPERSCRIPT for ease of comparison. Sink particles, representing stars, are depicted as circles, with their radius reflecting their mass. We note that both the gas and the dust exhibit similar large-scale structural features. For structural differences at smaller scales, refer to Figure 5.
Refer to caption
Figure 2: The rolling median of the accreted dust-to-gas (ADG) mass ratio μadgsuperscript𝜇adg\mu^{\rm adg}italic_μ start_POSTSUPERSCRIPT roman_adg end_POSTSUPERSCRIPT calculated for sink particles formed within a cloud of initial mass Mcloud=2×103Msubscript𝑀cloud2superscript103subscriptMdirect-productM_{\rm cloud}=2\times 10^{3}\rm M_{\odot}italic_M start_POSTSUBSCRIPT roman_cloud end_POSTSUBSCRIPT = 2 × 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT. The results are computed within logarithmic mass bins and normalised to the initial mean dust-to-gas (DTG) ratio of the cloud μ0dgdelimited-⟨⟩superscriptsubscript𝜇0dg\langle\mu_{0}^{\rm dg}\rangle⟨ italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_dg end_POSTSUPERSCRIPT ⟩. The darkly shaded and lightly shaded regions correspond to the interquartile and interdecile ranges respectively. The plot contrasts the full-physics runs with a run with limited dust physics, where m2e4_0.1 represents the full-physics fiducial run, m2e4_0.1_nomag corresponds to a run with neutral dust grains (no Lorentz forces), m2e4_0.1_norad corresponds to a run without radiation pressure forces on grains, and m2e4_0.1_pass represents a ’passive’ grain run with only drag forces and no back-reaction forces on the gas. Subsolar mass stars exhibit a scattered ADG ratio centered around the average value of μ0dgsimilar-toabsentdelimited-⟨⟩subscriptsuperscript𝜇dg0\sim\langle\mu^{\rm dg}_{0}\rangle∼ ⟨ italic_μ start_POSTSUPERSCRIPT roman_dg end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ⟩. The fiducial run reveals a clear trend of reduced ADG ratios at higher sink masses influenced by radiation forces on dust, which diminishes when radiative pressure forces are disabled.

While dust and gas exhibit coherence on cloud-size scales, stars primarily grow by accreting the local mixture of dust and gas, a process that determines various final stellar properties. Consequently, variations in the local DTG ratio can influence the properties of the final stellar population. Due to the simulation’s limited spatial resolution for sink accretion (10similar-toabsent10\sim 10∼ 10 AU), the fate of dust, whether it gets accreted by a star or is ejected, remains ambiguous. To address this uncertainty, we calculate a more robust measure: the accreted dust-to-gas (ADG) ratio, which represents the mass ratio of dust and gas elements satisfying accretion at the current resolution. We acknowledge that this measure has the potential for overestimation, considering the possibility of dust expulsion if accretion were tracked at smaller spatial scales.

3.2 Reduced Dust Accretion in Massive Stars

In Figure 2, we present the moving median ADG ratio for individual sink particles, calculated within logarithmic stellar mass bins, for a stellar population. We present the sinks formed with our larger cloud with a mass of Mcloud=2×104Msubscript𝑀cloud2superscript104subscriptMdirect-productM_{\rm cloud}=2\times 10^{4}\rm M_{\odot}italic_M start_POSTSUBSCRIPT roman_cloud end_POSTSUBSCRIPT = 2 × 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT to ensure a robust statistical dataset. To discern the contributions of various physical processes to the results, we compare the ADG ratios in our fiducial full-physics run with runs involving limited physics. Specifically, we experiment with runs, including a “passive” grain run where grains experience only drag forces, without Lorentz or radiation pressure forces and without exerting any back-reaction forces. We also examine a run in which grains experience both drag and Lorentz forces, and in turn induce back-reaction forces on the gas. However, in this specific run, the dust particles neither explicitly experience radiation forces nor is the dust opacity explicitly computed directly from the dust distribution. Instead, the opacity of the gas is calculated by assuming a constant DTG ratio. We also consider a scenario where we include all dust physics except Lorentz forces, i.e. the case of neutral dust grains. Finally, we contrast these scenarios with a comprehensive full-physics fiducial run, which encompasses all the aforementioned physical effects, along with explicit coupling of radiation to the dust grains.

As shown in the figure, we note a larger dispersions in the ADG ratio among sub-solar mass sinks for all simulation runs, with an average value converging to 1similar-toabsent1\sim 1∼ 1. This dispersion is primarily ascribed to the inherent limitations in resolution, set at around 102Msimilar-toabsentsuperscript102subscriptMdirect-product\sim 10^{-2}\,\rm M_{\odot}∼ 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, which results in low-mass sinks accreting a relatively limited number of gas and dust elements, and hence exhibiting greater stochastic noise.

For sinks of higher masses (28Mabsent28subscriptMdirect-product\geq 2-8\rm M_{\odot}≥ 2 - 8 roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT), a trend emerges: the ADG ratio decreases with increasing stellar mass. This phenomenon has implications for stellar properties and the abundance levels of elements commonly found in dust, including CNO, Si, Mg, and Fe, which we anticipate to be significantly reduced. These levels can plummet to nearly an order of magnitude lower than the original solid-phase abundances observed within the cloud. Strikingly, this trend becomes negligible when the dust opacity is not directly computed from the dust distribution, and the grains are not subject to radiation pressure forces. Furthermore, this pattern is virtually indiscernible when the grains are treated as passive entities within the system. Hence, we infer that this phenomenon is likely driven by radiation pressure which expels dust grains away from the vicinity of the sink particle. Massive stars, emitting intense radiation, exert substantial radiation pressure on the surrounding dust and gas, thereby significantly affecting their dynamics. This effect is expected to dominate the dynamics when radiation pressure surpasses gravitational forces, a condition expected to be met at specific stellar luminosities and/or masses.

In addition to the aforementioned tests, we conducted simulations to assess the impact of various stellar feedback mechanisms, including protostellar jets, stellar winds, and supernova feedback, by systematically enabling and disabling each in turn. We also explored the influence of magnetic fields through radiative hydrodynamical simulations, distinct from our tests involving neutral grains. We found that the identified trend remains robust across different choices of these model parameters.

3.3 Dust Evacuation

3.3.1 Toy Model

We can gain a simple understanding into whether a star would accrete a dust grain by examining the local dust dynamics through a set of simple approximations. Consider a scenario involving an inflowing shell of gas and dust influenced by a combination of drag and radiation from the star pushing the shell outward. The grains should reach a terminal velocity, where they drift relative to the gas at a velocity wssubscript𝑤𝑠w_{s}italic_w start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT. For simplicity, let us assume sub-sonic drift with vinflowϵcssimilar-tosubscript𝑣inflowitalic-ϵsubscript𝑐𝑠v_{\rm inflow}\sim\epsilon c_{s}italic_v start_POSTSUBSCRIPT roman_inflow end_POSTSUBSCRIPT ∼ italic_ϵ italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT where ϵitalic-ϵ\epsilonitalic_ϵ characterises the efficiency of gravitational in-fall with respect to the sound speed, and that the system is in the Rayleigh limit, characterised by Qextϵgrainproportional-tosubscriptdelimited-⟨⟩𝑄extsubscriptitalic-ϵgrain\langle Q\rangle_{\rm ext}\propto\epsilon_{\rm grain}⟨ italic_Q ⟩ start_POSTSUBSCRIPT roman_ext end_POSTSUBSCRIPT ∝ italic_ϵ start_POSTSUBSCRIPT roman_grain end_POSTSUBSCRIPT, allowing for a closed-form analytic solution. The amount of dust accreted by a star would then depend on the ratio of the terminal drift velocity of dust grains away from the star to their velocity through the in-falling gas vinflowsubscript𝑣inflowv_{\rm inflow}italic_v start_POSTSUBSCRIPT roman_inflow end_POSTSUBSCRIPT. This relationship can be approximated as follows:

wsvinflowsubscript𝑤ssubscript𝑣inflow\displaystyle\frac{w_{\rm s}}{v_{\rm inflow}}divide start_ARG italic_w start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT end_ARG start_ARG italic_v start_POSTSUBSCRIPT roman_inflow end_POSTSUBSCRIPT end_ARG aradtsϵcsQextF4πcϵρgcs2Qexturadϵuthermalsimilar-toabsentsubscript𝑎radsubscript𝑡𝑠italic-ϵsubscript𝑐𝑠similar-tosubscriptdelimited-⟨⟩𝑄extsubscript𝐹4𝜋𝑐italic-ϵsubscript𝜌𝑔superscriptsubscript𝑐𝑠2similar-tosubscriptdelimited-⟨⟩𝑄extsubscript𝑢raditalic-ϵsubscript𝑢thermal\displaystyle\sim\frac{a_{\rm rad}t_{s}}{\epsilon c_{s}}\sim\frac{\langle Q% \rangle_{\rm ext}F_{\star}}{4\pi c\epsilon\rho_{g}c_{s}^{2}}\sim\frac{\langle Q% \rangle_{\rm ext}\,u_{\rm rad}}{\epsilon u_{\rm thermal}}∼ divide start_ARG italic_a start_POSTSUBSCRIPT roman_rad end_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_ARG start_ARG italic_ϵ italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_ARG ∼ divide start_ARG ⟨ italic_Q ⟩ start_POSTSUBSCRIPT roman_ext end_POSTSUBSCRIPT italic_F start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG 4 italic_π italic_c italic_ϵ italic_ρ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ∼ divide start_ARG ⟨ italic_Q ⟩ start_POSTSUBSCRIPT roman_ext end_POSTSUBSCRIPT italic_u start_POSTSUBSCRIPT roman_rad end_POSTSUBSCRIPT end_ARG start_ARG italic_ϵ italic_u start_POSTSUBSCRIPT roman_thermal end_POSTSUBSCRIPT end_ARG (3)
QextL4πcr2ϵρgcs2similar-toabsentsubscriptdelimited-⟨⟩𝑄extsubscript𝐿4𝜋𝑐superscript𝑟2italic-ϵsubscript𝜌𝑔superscriptsubscript𝑐𝑠2\displaystyle\sim\frac{\langle Q\rangle_{\rm ext}L_{\star}}{4\pi cr^{2}% \epsilon\rho_{g}c_{s}^{2}}∼ divide start_ARG ⟨ italic_Q ⟩ start_POSTSUBSCRIPT roman_ext end_POSTSUBSCRIPT italic_L start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG 4 italic_π italic_c italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ϵ italic_ρ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG
0.1(Qext0.2)(LL)(0.01pcr)2(1ϵ)(103cm3nH)(20KT),similar-toabsent0.1subscriptdelimited-⟨⟩𝑄ext0.2subscript𝐿subscriptLdirect-productsuperscript0.01pc𝑟21italic-ϵsuperscript103superscriptcm3subscript𝑛H20K𝑇\displaystyle\sim 0.1\left(\frac{\langle Q\rangle_{\rm ext}}{0.2}\right)\left(% \frac{L_{\star}}{\rm L_{\odot}}\right)\left(\frac{0.01\,\rm pc}{r}\right)^{2}% \left(\frac{1}{\epsilon}\right)\left(\frac{10^{3}\,\rm cm^{-3}}{n_{\rm H}}% \right)\left(\frac{20\rm K}{T}\right),∼ 0.1 ( divide start_ARG ⟨ italic_Q ⟩ start_POSTSUBSCRIPT roman_ext end_POSTSUBSCRIPT end_ARG start_ARG 0.2 end_ARG ) ( divide start_ARG italic_L start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG roman_L start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT end_ARG ) ( divide start_ARG 0.01 roman_pc end_ARG start_ARG italic_r end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( divide start_ARG 1 end_ARG start_ARG italic_ϵ end_ARG ) ( divide start_ARG 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT end_ARG start_ARG italic_n start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT end_ARG ) ( divide start_ARG 20 roman_K end_ARG start_ARG italic_T end_ARG ) ,

where F=L/r2subscript𝐹subscript𝐿superscript𝑟2F_{\star}=L_{\star}/r^{2}italic_F start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT = italic_L start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT / italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT is the incident flux for a star of luminosity Lsubscript𝐿L_{\star}italic_L start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT on a grain at some radial distance r𝑟ritalic_r situated within a region of number density nHsubscript𝑛Hn_{\rm H}italic_n start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT. Assuming that LM3.5proportional-tosubscript𝐿superscriptsubscript𝑀3.5L_{\star}\propto M_{\star}^{3.5}italic_L start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT ∝ italic_M start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3.5 end_POSTSUPERSCRIPT, we determine that this ratio reaches unity when M2Msimilar-tosubscript𝑀2subscriptMdirect-productM_{\star}\sim 2\rm M_{\odot}italic_M start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT ∼ 2 roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT given the parameters we consider above.

We emphasize to the readers that this model is intentionally simplistic, designed to offer qualitative physical insight into the relevant parameters for this process. The model does not incorporate detailed models of accretion, density profiles of dust and/or gas, an accurate description of the object’s luminosity, or turbulence in the accretion flow, among other factors, all of which are expected to vary spatially and temporally as the star evolves. An additional point worth highlighting is that the ratio does not need to exceed 1 for the ADG ratio to be lower than the DTG ratio of the cloud; even modest values of outward dust drift relative to gas inflow would lead to reduced dust accretion.

We present a test of this hypothesis in Figure 3, where we use the full expression for tssubscript𝑡𝑠t_{s}italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT from Equation 2 to account for the subsonic and supersonic limits. On the left, we show the mean dust drift velocity, calculated as the mean radial velocity of dust relative to a nearby sink subtracted from the mean radial velocity of the gas computed for a discretized 3D grid. This is denoted as vdustvgaswsdelimited-⟨⟩subscript𝑣dustdelimited-⟨⟩subscript𝑣gassubscript𝑤𝑠\langle v_{\rm dust}\rangle-\langle v_{\rm gas}\rangle\equiv w_{s}⟨ italic_v start_POSTSUBSCRIPT roman_dust end_POSTSUBSCRIPT ⟩ - ⟨ italic_v start_POSTSUBSCRIPT roman_gas end_POSTSUBSCRIPT ⟩ ≡ italic_w start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, and the value is then normalized to the local sound speed at the position of the shell. Additionally, each data point is color-coded based on the mean stellar accretion rate, averaged over approximately 1 million years. We analyze the relationship between these parameters and the value of L/nHr2subscript𝐿subscript𝑛Hsuperscript𝑟2L_{\star}/n_{\rm H}r^{2}italic_L start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT / italic_n start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT within the grid element.

As expected from the simple relationship described in Equation 3 plotted in the dashed black line, the dust drift velocity is proportional to L/nHr2subscript𝐿subscript𝑛Hsuperscript𝑟2L_{\star}/n_{\rm H}r^{2}italic_L start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT / italic_n start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, maintaining this proportionality until ws/cs1similar-tosubscript𝑤𝑠subscript𝑐𝑠1w_{s}/c_{s}\sim 1italic_w start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT / italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ∼ 1. Beyond this point, as the flow transitions to the supersonic regime, the correlation transitions to a square root relationship with L/nHr2subscript𝐿subscript𝑛Hsuperscript𝑟2L_{\star}/n_{\rm H}r^{2}italic_L start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT / italic_n start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. We extend this assessment by examining the relationship between mean dust drift velocity and the mass of the central sink particle. Similarly, wssubscript𝑤𝑠w_{s}italic_w start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT positively correlates with the star’s mass, or in other words, with the time-integrated accretion rate, which, in turn, is associated with the integrated luminosity of the sink. Additionally, we observe a correlation with the time-averaged sink accretion rate, where periods of high accretion are associated with elevated dust drift velocities and a higher value of L/nHr2subscript𝐿subscript𝑛Hsuperscript𝑟2L_{\star}/n_{\rm H}r^{2}italic_L start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT / italic_n start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. This correlation is expected since it corresponds to increased luminosity and a higher gas density in the vicinity of the star.

Refer to caption
Figure 3: Left: The mean dust drift velocity ws=vdustvgasdelimited-⟨⟩subscript𝑤𝑠delimited-⟨⟩subscript𝑣dustdelimited-⟨⟩subscript𝑣gas\langle w_{s}\rangle=\langle v_{\rm dust}\rangle-\langle v_{\rm gas}\rangle⟨ italic_w start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ⟩ = ⟨ italic_v start_POSTSUBSCRIPT roman_dust end_POSTSUBSCRIPT ⟩ - ⟨ italic_v start_POSTSUBSCRIPT roman_gas end_POSTSUBSCRIPT ⟩ normalized to the sound speed cssubscript𝑐𝑠c_{s}italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT around sink particles with final M2Mgreater-than-or-equivalent-tosubscript𝑀2subscript𝑀direct-productM_{\star}\gtrsim 2M_{\odot}italic_M start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT ≳ 2 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT across time. Each data point, color-coded to represent the time-averaged mean accretion rate over a 1 Myr period, corresponds to mean values measured within a discretized 3D grid centered around the sinks. We plot vdustvgasdelimited-⟨⟩subscript𝑣dustdelimited-⟨⟩subscript𝑣gas\langle v_{\rm dust}\rangle-\langle v_{\rm gas}\rangle⟨ italic_v start_POSTSUBSCRIPT roman_dust end_POSTSUBSCRIPT ⟩ - ⟨ italic_v start_POSTSUBSCRIPT roman_gas end_POSTSUBSCRIPT ⟩, against the luminosity of the sink particle in its proximity, divided by its gas number density and the square of its radial distance from the sink, L/nHr2subscript𝐿subscript𝑛Hsuperscript𝑟2L_{\star}/n_{\rm H}r^{2}italic_L start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT / italic_n start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. The dashed black line represents the theoretical prediction from Equation 3, incorporating both subsonic (wsL/nHr2)proportional-todelimited-⟨⟩subscript𝑤𝑠subscript𝐿subscript𝑛Hsuperscript𝑟2\left(\langle w_{s}\rangle\propto L_{\star}/n_{\rm H}r^{2}\right)( ⟨ italic_w start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ⟩ ∝ italic_L start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT / italic_n start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) and supersonic regimes (wsL/nHr2)proportional-todelimited-⟨⟩subscript𝑤𝑠subscript𝐿subscript𝑛Hsuperscript𝑟2\left(\langle w_{s}\rangle\propto\sqrt{L_{\star}/n_{\rm H}r^{2}}\right)( ⟨ italic_w start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ⟩ ∝ square-root start_ARG italic_L start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT / italic_n start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ). The dust drift exhibits a positive correlation with L/nHr2subscript𝐿subscript𝑛Hsuperscript𝑟2L_{\star}/n_{\rm H}r^{2}italic_L start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT / italic_n start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, as theoretically predicted. There is scatter away from the theoretical prediction due to the model’s simplicity and the influence of other forces and turbulence on the dynamics. Right: The mean drift velocity against the mass of the sink particle. Note that the discreteness in stellar mass corresponds to the simulation resolution. The dust drift positively correlates with the sink’s mass and mean accretion rate, in agreement with predictions.

3.3.2 A Case Study

In Figure 4, we present a case study of this phenomenon, focusing on a region around a selected sink particle. We chose this particular candidate after examining several others, as it demonstrates the median behaviour with minimal noise. We also carefully chose specific timesteps to minimise the impact of noise on the data and to clarify the dust evacuation process. Each line in the plot corresponds to a snapshot in time and is colour-coded accordingly. All values are computed as the mean values within narrow radial shells centered around the star. The top four plots represent the state of the system during the initial creation of the dust-evacuated region, while the lower section corresponds to the subsequent dispersal of the dust-evacuated region.

At early times, there is only a minimal reduction in the DTG near the sink particle. For instance, at 104superscript10410^{-4}10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT pc from the sink, the DTG ratio is only half of the average initial value within the cloud. As we move farther from the sink, we observe an accumulation of dust, leading to a DTG ratio of approximately μdg/μ0dg2similar-tosuperscript𝜇dgsubscriptsuperscript𝜇dg02\mu^{\rm dg}/\mu^{\rm dg}_{0}\sim 2italic_μ start_POSTSUPERSCRIPT roman_dg end_POSTSUPERSCRIPT / italic_μ start_POSTSUPERSCRIPT roman_dg end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∼ 2, which gradually approaches the mean value as we reach a distance of about 0.1 parsecs from the sink.

As the sink continues to accrete matter and increase in mass, we observe corresponding changes in the gas environment. The gas number density increases from its peak value of 107cm3superscript107superscriptcm310^{7}\rm cm^{-3}10 start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT to 109cm3superscript109superscriptcm310^{9}\rm cm^{-3}10 start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT, and the gas temperature transitions from an initial value of T20similar-to𝑇20T\sim 20italic_T ∼ 20K to T100similar-to𝑇100T\sim 100italic_T ∼ 100K. As material accretes onto the sink, the dust is entrained alongside the gas, and the peak in the DTG ratio shifts closer to the sink. Additionally, we observe that the peak DTG ratio increases. This effect is primarily attributed to the gas density drop** more rapidly than the dust density in that specific region. We propose that this effect is driven by radiation pressure forces that induce a net outward motion of the dust, in contrast to the net inflow of gas. This process gives rise to the formation of a dust-evacuated region surrounding the star. As a consequence, a dust shell, approximately 102superscript10210^{-2}10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT pc in thickness, emerges, where μ0dg/μ0dg5similar-todelimited-⟨⟩subscriptsuperscript𝜇dg0delimited-⟨⟩subscriptsuperscript𝜇dg05\langle\mu^{\rm dg}_{0}\rangle/\langle\mu^{\rm dg}_{0}\rangle\sim 5⟨ italic_μ start_POSTSUPERSCRIPT roman_dg end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ⟩ / ⟨ italic_μ start_POSTSUPERSCRIPT roman_dg end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ⟩ ∼ 5. In the immediate vicinity of the star, this effect is accompanied by values drop** as low as μdg/μ0dg102similar-todelimited-⟨⟩superscript𝜇dgsubscriptsuperscript𝜇dg0superscript102\langle\mu^{\rm dg}\rangle/\mu^{\rm dg}_{0}\sim 10^{-2}⟨ italic_μ start_POSTSUPERSCRIPT roman_dg end_POSTSUPERSCRIPT ⟩ / italic_μ start_POSTSUPERSCRIPT roman_dg end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∼ 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT in the central region.

In the four bottom plots, we present the same parameters during a phase characterised by a decline in the accretion rate, as indicated by the sink mass-time plot. We observe a reduction in the gas number density and temperature over time, consistent with expectations during a reduced accretion phase. Additionally, we note a gradual increase in the DTG ratio within the innermost regions, attributable to the diminishing gas density near the central region. The reduced gas number density also leads to weaker drag forces experienced by the dust grains, resulting in the outward shift of the peak DTG ratio and the reduction of its maximum value as the dust shell disperses.

Refer to caption
Refer to caption
Figure 4: The temporal evolution of environmental properties detailing the creation and subsequent dispersion of a dust-evacuated region around a reference sink particle. We only show specific timesteps that most clearly highlight the evolution of this process. Each line in the plot corresponds to a given timestep and is color-coded accordingly. The upper four plots, in the hot colormap, depict the formation of a dust-evacuated region during a high accretion state, while the lower four, in the cold colormap, show its dispersal during reduced accretion. The properties are evaluated within narrow radial shells. Top Left: The mean dust-to-gas (DTG) μdg(r)delimited-⟨⟩superscript𝜇dg𝑟\langle\mu^{\rm dg}(r)\rangle⟨ italic_μ start_POSTSUPERSCRIPT roman_dg end_POSTSUPERSCRIPT ( italic_r ) ⟩ ratio normalized to the initial mean value of the cloud μ0dgdelimited-⟨⟩subscriptsuperscript𝜇dg0\langle\mu^{\rm dg}_{0}\rangle⟨ italic_μ start_POSTSUPERSCRIPT roman_dg end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ⟩. Top Right: The mean number density of the gas nH(r)subscript𝑛H𝑟n_{\rm H}(r)italic_n start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT ( italic_r ). Bottom Left: The mean gas temperature T(r)𝑇𝑟T(r)italic_T ( italic_r ). Bottom Right: The sink particle’s mass evolution. The plots illustrate the in-fall of material towards the sink, raising the central gas density. Ongoing accretion and heightened luminosity increases the gas temperature due to radiation. As gas flows inward, it carries dust closer, but the enhanced luminosity counteracts the motion, propelling dust outward to form a 102similar-toabsentsuperscript102\sim 10^{-2}∼ 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT pc thick dustshell, with a peak DTG ratio of μdg5μ0dgsimilar-todelimited-⟨⟩superscript𝜇dg5delimited-⟨⟩subscriptsuperscript𝜇dg0\langle\mu^{\rm dg}\rangle\sim 5\langle\mu^{\rm dg}_{0}\rangle⟨ italic_μ start_POSTSUPERSCRIPT roman_dg end_POSTSUPERSCRIPT ⟩ ∼ 5 ⟨ italic_μ start_POSTSUPERSCRIPT roman_dg end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ⟩. As accretion slows down, the gas density and temperature decrease, while the DTG ratio rises in the inner regions, and the peak shifts outward. The plots illustrate variations in μdgsuperscript𝜇dg\mu^{\rm dg}italic_μ start_POSTSUPERSCRIPT roman_dg end_POSTSUPERSCRIPT around a sink particle induced by radiation pressure, and highlight the development of a dust-evacuated zone encircled by a dust-rich shell.

While we demonstrated that both dust and gas exhibit similar features on large scales, we will now discuss the differences that arise on smaller scales. In Figure 5, we transition from the overview of the gas surface density of the entire cloud in our m2e3_0.1_hires run (6pc×6pc×6pc6pc6pc6pc6\text{pc}\times 6\text{pc}\times 6\text{pc}6 pc × 6 pc × 6 pc) to a reduced volume. Specifically, the middle panel focuses on a smaller volume with dimensions of (0.01pc×0.01pc×0.005pc0.01pc0.01pc0.005pc0.01\text{pc}\times 0.01\text{pc}\times 0.005\text{pc}0.01 pc × 0.01 pc × 0.005 pc), presenting the integrated 2D surface density of the gas through a slice in that region. Individual dust particles, color-coded to denote their respective grain sizes, are overlaid on the gas distribution. The right panel zooms in on an even more compact space with dimensions of (0.005pc×0.005pc×0.0025pc0.005pc0.005pc0.0025pc0.005\text{pc}\times 0.005\text{pc}\,\times 0.0025\text{pc}0.005 pc × 0.005 pc × 0.0025 pc), showcasing the mean DTG ratio within the volume. Both zoomed-in plots are centered around a specifically chosen sink particle. It is worth noting that the selection of this particular sink particle and snapshot is intentional, as they effectively illustrate the dust evacuation and dust pile-up phenomena.

The plot highlights a key observation: on sub-parsec scales, particularly in the proximity of stars, the spatial distribution of dust particles deviates from that of the gas. Further as indicated by the plots in Figure 4, we observe a dust-suppressed zone near the sink particle followed by dust-rich region, indicative of the presence of a dust pile-up or a dust “shell” envelo** these stellar objects.

Refer to caption
Figure 5: The 2D integrated gas surface density of the m2e3_0.1_hires simulation. Left: The gas surface density of the cloud, with white circles representing sink particles, sized according to their mass. Middle: The gas surface density centered around a specific sink particle, with individual dust particles color-coded by grain size. Right: The dust-to-gas mass ratio (DTG) normalized to the mean DTG of the cloud within a 0.01 pc region centered around the sink particle. This figure illustrates the presence of a dust-evacuated zone around a sink particle, featuring a dust pile-up or “shell” surrounding the sink.
Refer to caption
Figure 6: The rolling median of the accreted dust-to-gas (ADG) ratio μadgsuperscript𝜇adg\mu^{\rm adg}italic_μ start_POSTSUPERSCRIPT roman_adg end_POSTSUPERSCRIPT normalized to the initial mean DTG ratio of the cloud μ0dgdelimited-⟨⟩superscriptsubscript𝜇0dg\langle\mu_{0}^{\rm dg}\rangle⟨ italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_dg end_POSTSUPERSCRIPT ⟩ for clouds with Mcloud2×103Msimilar-tosubscript𝑀cloud2superscript103subscriptMdirect-productM_{\rm cloud}\sim 2\times 10^{3}\rm M_{\odot}italic_M start_POSTSUBSCRIPT roman_cloud end_POSTSUBSCRIPT ∼ 2 × 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT and Mcloud2×104Msimilar-tosubscript𝑀cloud2superscript104subscriptMdirect-productM_{\rm cloud}\sim 2\times 10^{4}\rm M_{\odot}italic_M start_POSTSUBSCRIPT roman_cloud end_POSTSUBSCRIPT ∼ 2 × 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT. We show both clouds simulated at a resolution of Δm102Msimilar-toΔ𝑚superscript102subscriptMdirect-product\Delta m\sim 10^{-2}\rm M_{\odot}roman_Δ italic_m ∼ 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT. Darkly shaded and lightly shaded regions indicate the interquartile and interdecile percentile ranges, respectively. The relatively consistent trend suggests that, within the mass range investigated here, the cloud mass does not affect the trend.
Refer to caption
Figure 7: The normalized rolling median of the accreted dust-to-gas ratio (μadgsuperscript𝜇adg\mu^{\rm adg}italic_μ start_POSTSUPERSCRIPT roman_adg end_POSTSUPERSCRIPT) in simulations with different resolutions. Shading indicates interquartile and interdecile ranges. Dark blue line shows a 2×103M2superscript103subscriptMdirect-product2\times 10^{3}\rm M_{\odot}2 × 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT cloud at resolutions of Δm103similar-toΔ𝑚superscript103\Delta m\sim 10^{-3}roman_Δ italic_m ∼ 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT while the pink line shows the same cloud at Δm102Msimilar-toΔ𝑚superscript102subscriptMdirect-product\Delta m\sim 10^{-2}\rm M_{\odot}roman_Δ italic_m ∼ 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT. Higher resolution reduces scatter, and aligns μadgsuperscript𝜇adg\mu^{\rm adg}italic_μ start_POSTSUPERSCRIPT roman_adg end_POSTSUPERSCRIPT closer to μ0dgdelimited-⟨⟩superscriptsubscript𝜇0dg\langle\mu_{0}^{\rm dg}\rangle⟨ italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_dg end_POSTSUPERSCRIPT ⟩ before declining at higher sink masses.

3.4 Effects of Cloud Mass

We consider how this phenomenon is affected by cloud properties. In Figure 6, we compare the ADG ratio in two distinct clouds, both simulated with our full physics setup. Both clouds are simulated with a common resolution of Δm102Msimilar-toΔ𝑚superscript102subscriptMdirect-product\Delta m\sim 10^{-2}\rm M_{\odot}roman_Δ italic_m ∼ 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, with one having a mass of 2×103M2superscript103subscriptMdirect-product2\times 10^{3}\rm M_{\odot}2 × 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT and the other 2×104M2superscript104subscriptMdirect-product2\times 10^{4}\rm M_{\odot}2 × 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT The results obtained reveal a consistent trend in the ADG ratio, which remains unaffected by variations in cloud mass. Across this mass range, the influence of radiation pressure, inherently a localised phenomenon, exerts comparable effects across diverse cloud masses. Nevertheless, although this phenomenon primarily impacts the ADG ratio of a specific star, its influence could extend to larger radii. As more massive stars form, their high luminosities could also lead to a reduction in dust content for neighbouring stars.

While the primary emphasis of this work centers on reduced ADG ratios for high-mass stars, our simulations also unveil a subset of low-mass stars that are rich in dust content. Specifically, within the subset of solar and subsolar mass stars, we observe instances of heightened ADG ratios. However, we refrain from attributing this phenomenon to a single cause. Instead, we postulate that it likely stems from the intricate interplay between turbulent motions and radiation pressure emanating from neighboring stars, leading to the redistribution of dust. Hence, stars emerging from regions with dust over-densities are prone to experiencing increased levels of dust accretion.

3.5 Effects of Simulation Resolution

In Figure 7, we present a comparative analysis of the ADG ratios within stellar populations formed in a cloud with a mass of Mcloud2×103Msimilar-tosubscript𝑀cloud2superscript103subscriptMdirect-productM_{\rm cloud}\sim 2\times 10^{3}\rm M_{\odot}italic_M start_POSTSUBSCRIPT roman_cloud end_POSTSUBSCRIPT ∼ 2 × 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, utilising simulations at different resolutions (Δm103similar-toΔ𝑚superscript103\Delta m\sim 10^{-3}roman_Δ italic_m ∼ 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT and Δm102Msimilar-toΔ𝑚superscript102subscriptMdirect-product\Delta m\sim 10^{-2}\rm M_{\odot}roman_Δ italic_m ∼ 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT). As anticipated, our observations indicate that increasing the simulation resolution decreases the scatter in the distribution. In addition, for the higher resolution cloud, the ADG ratio tends to converge to μ0dgsubscriptsuperscript𝜇dg0\mu^{\rm dg}_{0}italic_μ start_POSTSUPERSCRIPT roman_dg end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT for relatively low-mass stars (those with masses less than a few solar masses). This aligns with the expectation that stars with lower luminosities lack the necessary conditions to evacuate dust, and therefore would have ADG ratios that mirror the average DTG ratio within the cloud. These improvements can be attributed to the enhanced sampling capabilities, allowing for a more precise resolution of accretion and sink formation.

However, as shown in the plot, our findings are sensitive to resolution. This observation is unsurprising given the inherent challenges in accurately modelling accretion around luminous stars (Krumholz et al., 2009; Krumholz & Thompson, 2012; Rosen et al., 2016). Achieving an accurate depiction of relative dust-to-gas accretion onto a star, accounting for associated radiation effects, requires significantly higher resolutions than currently feasible with our simulations.

To illustrate, considering a 1 solar mass star that would have accreted approximately 0.01Msimilar-toabsent0.01subscript𝑀direct-product\sim 0.01M_{\odot}∼ 0.01 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT of dust mass, assuming μdg102similar-tosuperscript𝜇dgsuperscript102\mu^{\rm dg}\sim 10^{-2}italic_μ start_POSTSUPERSCRIPT roman_dg end_POSTSUPERSCRIPT ∼ 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT, at the 102superscript10210^{-2}10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT resolution (bearing in mind that the dust is up-sampled by a ratio of 4 times the gas resolution), this roughly corresponds accretion of similar-to\sim 4 dust particles. Consequently, we do not necessarily anticipate convergent results at our low resolutions, particularly at low sink masses. However, this approach still provides valuable insights when comparing simulations conducted at the same resolution. We hope that our results motivate detailed zoom-in simulations of dusty accretion around protostellar objects and onto massive stars, to provide a more detailed study of the dust evacuation phenomenon.

To evaluate the effects of resolution on our findings, we conducted a series of idealized tests of singular Shu (1977) collapse scenarios at resolutions of 102superscript10210^{-2}10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT, 103superscript10310^{-3}10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT, and 104Msuperscript104subscriptMdirect-product10^{-4}\rm M_{\odot}10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT. These tests track the problem hydrodynamically, allowing sink particle formation with subsequent accretion and radiation. To simplify our analysis, we exclude the effects of magnetic fields and protostellar jets and stellar winds. We find that the observed phenomena qualitatively persist across different resolutions; however, convergence is not achieved. In particular, neither the spatial extent of the dust evacuation zone nor the precise magnitude of the stellar ADG ratio converge with increasing resolution.

Considering the complexities of radiation-limited accretion, it remains uncertain whether the system, as modeled in our simulations, should converge within our resolution range. Higher resolution simulations capture finer-scale structures, unveiling more intricate dynamics around the sink which thereby altering the magnitude and scale of dust evacuation. Additionally, the mechanisms of dust and gas accretion, whether via spherical collapse or disk accretion, and their dynamics are resolution-dependent and influenced by the specific physics in our simulations.

3.6 Effects of Altered Grain Size Distributions

Refer to caption
Figure 8: The rolling median of the accreted dust-to-gas (ADG) ratio (μadgsuperscript𝜇adg\mu^{\rm adg}italic_μ start_POSTSUPERSCRIPT roman_adg end_POSTSUPERSCRIPT) for sink particles formed within a cloud of initial mass Mcloud=2×103Msubscript𝑀cloud2superscript103subscriptMdirect-productM_{\rm cloud}=2\times 10^{3}\rm M_{\odot}italic_M start_POSTSUBSCRIPT roman_cloud end_POSTSUBSCRIPT = 2 × 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT (top), Mcloud=2×104Msubscript𝑀cloud2superscript104subscriptMdirect-productM_{\rm cloud}=2\times 10^{4}\rm M_{\odot}italic_M start_POSTSUBSCRIPT roman_cloud end_POSTSUBSCRIPT = 2 × 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT (bottom) with different initial grain-size distributions (ϵgrainmax=0.1μmsuperscriptsubscriptitalic-ϵgrainmax0.1𝜇m\epsilon_{\rm grain}^{\rm max}=0.1\rm\mu mitalic_ϵ start_POSTSUBSCRIPT roman_grain end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_max end_POSTSUPERSCRIPT = 0.1 italic_μ roman_m, ϵgrainmax=1μmsuperscriptsubscriptitalic-ϵgrainmax1𝜇m\epsilon_{\rm grain}^{\rm max}=1\rm\mu mitalic_ϵ start_POSTSUBSCRIPT roman_grain end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_max end_POSTSUPERSCRIPT = 1 italic_μ roman_m, and ϵgrainmax=10μmsuperscriptsubscriptitalic-ϵgrainmax10𝜇m\epsilon_{\rm grain}^{\rm max}=10\rm\mu mitalic_ϵ start_POSTSUBSCRIPT roman_grain end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_max end_POSTSUPERSCRIPT = 10 italic_μ roman_m). The darkly shaded and lightly shaded regions represent the interquartile and interdecile percentile ranges, respectively. The m2e4_10 run exhibits notable scatter and a low count of formed sink particles. To account for this, we scatter the ADG ratio for each sink particle. Refer to Section 3.5 for a discussion on potential drivers of this scatter. Changes in the grain size do not drive significant deviations from the reduced μadgsuperscript𝜇adg\mu^{\rm adg}italic_μ start_POSTSUPERSCRIPT roman_adg end_POSTSUPERSCRIPT trend for high mass stars, but larger grain sizes lead to fewer high-mass stars due to reduced star formation efficiency.

An additional variable to consider is the size of the dust grains. At extremely small grain sizes, the grains should closely trace the dynamics of the gas, whereas at larger grain sizes, the grains will be entirely decoupled from gas dynamics. In Figure 8, we explore the impact of various plausible grain sizes within the cloud environment, incorporating grains with maximum sizes of ϵgrainmax10μmsimilar-tosuperscriptsubscriptitalic-ϵgrainmax10𝜇m\epsilon_{\rm grain}^{\rm max}\sim 10\rm\mu mitalic_ϵ start_POSTSUBSCRIPT roman_grain end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_max end_POSTSUPERSCRIPT ∼ 10 italic_μ roman_m and ϵgrainmax1μmsimilar-tosuperscriptsubscriptitalic-ϵgrainmax1𝜇m\epsilon_{\rm grain}^{\rm max}\sim 1\rm\mu mitalic_ϵ start_POSTSUBSCRIPT roman_grain end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_max end_POSTSUPERSCRIPT ∼ 1 italic_μ roman_m, alongside our baseline assumption of ϵgrainmax0.1μmsimilar-tosuperscriptsubscriptitalic-ϵgrainmax0.1𝜇m\epsilon_{\rm grain}^{\rm max}\sim 0.1\rm\mu mitalic_ϵ start_POSTSUBSCRIPT roman_grain end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_max end_POSTSUPERSCRIPT ∼ 0.1 italic_μ roman_m. This is conducted while ensuring a fixed total dust mass within the simulation and maintaining a dynamic range of 100similar-toabsent100\sim 100∼ 100 in grain sizes. We show the results for our higher resolution Mcloud2×103Msimilar-tosubscript𝑀cloud2superscript103subscriptMdirect-productM_{\rm cloud}\sim 2\times 10^{3}\rm M_{\odot}italic_M start_POSTSUBSCRIPT roman_cloud end_POSTSUBSCRIPT ∼ 2 × 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT simulations, in addition to a Mcloud2×104Msimilar-tosubscript𝑀cloud2superscript104subscriptMdirect-productM_{\rm cloud}\sim 2\times 10^{4}\rm M_{\odot}italic_M start_POSTSUBSCRIPT roman_cloud end_POSTSUBSCRIPT ∼ 2 × 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT cloud at lower resolution. While in principle, grain size could influence our results, our findings for these more reasonable sizes are consistent with Equation 3, which demonstrates no grain-size dependence. It is worth noting that the grain size does have an impact on the total number of sink particles that form, with larger grains leading to a reduced overall sink count. This effect and related changes will be studied in detail in Soliman & Hopkins (in prep.).

However, we begin to see some deviations from the reported trend for our largest grain simulations. Note that we present individual μadgsuperscript𝜇adg\mu^{\rm adg}italic_μ start_POSTSUPERSCRIPT roman_adg end_POSTSUPERSCRIPT values for each sink particle in the simulation with Mcloud2×104Msimilar-tosubscript𝑀cloud2superscript104subscriptMdirect-productM_{\rm cloud}\sim 2\times 10^{4}\rm M_{\odot}italic_M start_POSTSUBSCRIPT roman_cloud end_POSTSUBSCRIPT ∼ 2 × 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT and ϵgrainmax=10μmsuperscriptsubscriptitalic-ϵgrainmax10𝜇m\epsilon_{\rm grain}^{\rm max}=10\rm\mu mitalic_ϵ start_POSTSUBSCRIPT roman_grain end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_max end_POSTSUPERSCRIPT = 10 italic_μ roman_m, opting against plotting the median due to a limited number of sink particles and a substantial dispersion in their μadgsuperscript𝜇adg\mu^{\rm adg}italic_μ start_POSTSUPERSCRIPT roman_adg end_POSTSUPERSCRIPT. The dispersion is likely influenced by several factors, including the relatively lower resolution. However, we posit that the increased grain size inherently contributes to a higher level of scatter. Recall that these simulation assume a fixed a grain size spectrum, where individual grain particles maintain a fixed size throughout the simulation. Consequently, as grain size increases, the individual grain count decreases, while each grain becomes more massive. This suggests that whether a particle undergoes accretion or expulsion has a more pronounced impact on the overall dust mass accreted by a sink, especially when compared to its smaller size/mass grain counterpart. Further, Equation 3 assumes that Qextsubscriptdelimited-⟨⟩𝑄ext\langle Q\rangle_{\rm ext}⟨ italic_Q ⟩ start_POSTSUBSCRIPT roman_ext end_POSTSUBSCRIPT is proportional to 2πϵgrain/λ2𝜋subscriptitalic-ϵgrain𝜆2\pi\epsilon_{\rm grain}/\lambda2 italic_π italic_ϵ start_POSTSUBSCRIPT roman_grain end_POSTSUBSCRIPT / italic_λ, where λ𝜆\lambdaitalic_λ denotes the radiation wavelength. However, for grains larger than 2π/λ2𝜋𝜆2\pi/\lambda2 italic_π / italic_λ, Qextsubscriptdelimited-⟨⟩𝑄ext\langle Q\rangle_{\rm ext}⟨ italic_Q ⟩ start_POSTSUBSCRIPT roman_ext end_POSTSUBSCRIPT approaches 1. As a result, wsϵgrain1proportional-tosubscript𝑤𝑠superscriptsubscriptitalic-ϵgrain1w_{s}\propto\epsilon_{\rm grain}^{-1}italic_w start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ∝ italic_ϵ start_POSTSUBSCRIPT roman_grain end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, introducing an inverse dependence on grain size into Equation 3. Considering the case of clouds with ϵgrainmax=10μmsuperscriptsubscriptitalic-ϵgrainmax10𝜇m\epsilon_{\rm grain}^{\rm max}=10\rm\mu mitalic_ϵ start_POSTSUBSCRIPT roman_grain end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_max end_POSTSUPERSCRIPT = 10 italic_μ roman_m, all grains fall within this regime for radiation with λ600μmless-than-or-similar-to𝜆600𝜇m\lambda\lesssim 600\rm\mu mitalic_λ ≲ 600 italic_μ roman_m, and larger grains fall into this regime for even shorter wavelengths. This wavelength range corresponds to the optical/UV range, where stars—especially more massive and hotter ones—typically emit peak radiation. Therefore, we would expect to see increases in μadgsuperscript𝜇adg\mu^{\rm adg}italic_μ start_POSTSUPERSCRIPT roman_adg end_POSTSUPERSCRIPT for such massive grains, although such large grains are not expected to be abundant in most GMCs.

In reality, dust grains undergo processes such as coagulation, accretion, sputtering, photodestruction, and shattering which result in both changes in the grain size distribution and the overall dust mass. These mechanisms, currently not captured in our simulations, could impact the interpretation of our results. For instance, if dust sublimation proceeds efficiently and the elements comprising the grains become well-mixed in the gas phase near the stars, they may be incorporated into the stars, potentially erasing the effects of dust evacuation on stellar abundances. However, the dust-evacuated zones around the stars would still persist or undergo further evacuation. We acknowledge these limitations and plan to address them in future work.

In Figure 9, we present the bivariate distribution showing the correlation between the local 3D gas density (ρgsubscript𝜌𝑔\rho_{g}italic_ρ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT) and the dust density (ρdsubscript𝜌𝑑\rho_{d}italic_ρ start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT). This distribution is weighted by the dust mass and computed at a spatial resolution of approximately 102superscript10210^{-2}10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT pc for the different grain size runs (ϵgrainmax=0.1,1,10,μmsubscriptsuperscriptitalic-ϵmaxgrain0.1110𝜇m\epsilon^{\rm max}_{\rm grain}=0.1,1,10,\mu\text{m}italic_ϵ start_POSTSUPERSCRIPT roman_max end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_grain end_POSTSUBSCRIPT = 0.1 , 1 , 10 , italic_μ m) at a time of approximately 2tdyn2subscript𝑡dyn2\,t_{\rm dyn}2 italic_t start_POSTSUBSCRIPT roman_dyn end_POSTSUBSCRIPT. For the smallest grains, the distribution exhibits a relatively narrow distribution, primarily centered around the theoretically predicted perfect coupling line ρd=μdgρgsubscript𝜌𝑑superscript𝜇dgsubscript𝜌𝑔\rho_{d}=\mu^{\rm dg}\rho_{g}italic_ρ start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT = italic_μ start_POSTSUPERSCRIPT roman_dg end_POSTSUPERSCRIPT italic_ρ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT. However, as the grain size increases, the distribution broadens, signifying a decreased level of coupling between the gas and dust components. Nevertheless, on average, the fluid remains sufficiently well-coupled to prevent large fluctuations in the DTG ratio.

Refer to caption
Figure 9: Bivariate distribution of the local 3D gas density (ρgsubscript𝜌𝑔\rho_{g}italic_ρ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT) and dust density (ρdsubscript𝜌𝑑\rho_{d}italic_ρ start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT), weighted by dust mass, illustrating the probability distribution around a grain at the resolution scale of approximately 102superscript10210^{-2}10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT pc. From top to bottom, we show the distribution for a run with ϵgrainmax=0.1,1,10μmsuperscriptsubscriptitalic-ϵgrainmax0.1110𝜇m\epsilon_{\rm grain}^{\rm max}=0.1,1,10\,\rm\mu mitalic_ϵ start_POSTSUBSCRIPT roman_grain end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_max end_POSTSUPERSCRIPT = 0.1 , 1 , 10 italic_μ roman_m respectively. We show the 1σ1𝜎1-\sigma1 - italic_σ (green), 2σ2𝜎2-\sigma2 - italic_σ (navy), 3σ3𝜎3-\sigma3 - italic_σ (orange), 4σ4𝜎4-\sigma4 - italic_σ (plum) contours. The diagonal dotted lines represent perfect dust-gas coupling (ρd=μdgρgsubscript𝜌𝑑superscript𝜇dgsubscript𝜌𝑔\rho_{d}=\mu^{\rm dg}\rho_{g}italic_ρ start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT = italic_μ start_POSTSUPERSCRIPT roman_dg end_POSTSUPERSCRIPT italic_ρ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT), while the horizontal line denotes uniform dust density. The distribution becomes progressively broader indicative of weaker coupling for larger grain sizes.

4 Discussion

4.1 Implications

Our findings, which reveal a diminished accretion of dust onto massive stars under the influence of radiation pressure, bear notable implications for the stellar abundances of CNO, Mg, Fe, and Si, elements known to preferentially deplete onto dust grains. This phenomenon was previously suggested in the literature as a potential explanation for the anomalous abundance ratios observed in nearby stars (Melendez et al., 2009; Gustafsson, 2018a, b; Mathews, 1967; Cochran & Ostriker, 1977). A distinctive feature of this model is its localized effect, with the evacuation zone extending up to 103similar-toabsentsuperscript103\sim 10^{-3}∼ 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT pc or a few hundred AUs. Consequently, this preferential accretion mechanism may contribute to elucidating the anomalous abundance ratios observed in specific nearby stars and the variations in abundance ratios among stellar pairs (Maia et al., 2014; Nissen et al., 2017; Ramírez et al., 2015; Biazzo et al., 2015; Teske et al., 2016a, b; Saffe et al., 2017; Oh et al., 2018, e.g.,).

Further, the observed decrease in dust content has significant implications for the gas dynamics and thermochemistry near the star. Lower dust content, associated with diminished opacities and cooling rates, would accelerate the expansion rate of HII regions (Ali, 2021). It would likely result in a larger ionized volume, owing to the reduced UV absorption by dust.

Moreover, protoplanetary disks would also potentially affected by this phenomenon. The DTG ratio plays a crucial role in determining the efficiency of dust radial drift, influencing the structural evolution of the disk (Toci et al., 2021). Moreover, as dust constitutes the fundamental building block of planets, diminished dust content is likely to influence the types of planets that can form and their compositions.

4.2 Caveats

The current study has inherent limitations that warrant careful consideration. A major constraint is the absence of a model for dust evolution; our approach relies on assuming a constant grain size distribution and a constant total dust mass throughout the simulation. Consequently, should the grain size distribution undergo significant changes, specific conclusions within our study may require reevaluation. In particular, if dust grains experience substantial destruction, whether through sublimation, shattering and/or sputtering, evacuated regions would likely persist or even intensify. However, if the elements released from the grains transition to the gas phase and continue to accrete onto stars, spatial variations in dust distribution might not necessarily indicate corresponding variations in stellar abundances.

Furthermore, our fiducial grain size distribution incorporates nanometer-sized grains. The existence of such small grains in these conditions is disputed, given the potential for coagulation or destruction due to the conditions in molecular clouds. Additionally, it is uncertain whether such grains can be modeled aerodynamically and whether they would adhere to the charge and mass scalings we assume. However, it is important to note that the results presented here predominantly depend on the grains holding the most mass, which, under an MRN spectrum, are the largest grains. Nevertheless, these considerations prompt us to explore different ranges of grain size in Section 3.6 to assess how varying grain properties might influence our results.

An additional constraint in our study is the resolution limit, which is discussed in more detail in Section 3.5. To ensure a statistically robust sample size for identifying the phenomenon of dust evacuation, we conducted global star formation simulations. However, the most interesting behavior occurs at the resolution limit of our study. Achieving detailed predictions for the implications of dust dynamics on stellar metallicities, protostellar envelope properties, and planet formation requires more intricate small-scale physics. Addressing these complexities realistically demands a substantial increase in resolution, which we plan to explore in future work.

5 Conclusions

This study introduces a set of RDMHD simulations of star forming GMCs as part of the STARFORGE project. These simulations encompass a detailed representation of individual star formation, accretion, and feedback mechanisms, while also explicitly considering the influence of the dynamics of dust grains. Our investigation focuses on the implications of these dynamics, specifically radiation-dust interactions, on the emergent properties of stellar populations.

Through our analysis, we find that when stars surpass a critical mass threshold (2Msimilar-toabsent2subscript𝑀direct-product\sim 2M_{\odot}∼ 2 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT), their luminosity exerts sufficient radiation pressure on neighbouring dust grains, ultimately resulting in their expulsion from the star’s accretion radius. This drives the formation of a dust-evacuated region of size 100similar-toabsent100\sim 100∼ 100 AU. Consequently, this process results in a mass-dependent adjustment in the ADG mass ratio incorporated into these stars via the accretion process. Furthermore, we investigate the potential implications of varying cloud mass, grain sizes, and other physical parameters such as decoupling the grains from magnetic fields and deactivating stellar winds and jets. However, our findings indicated that these variations had negligible effects within the parameter space of our study.

In summary, our findings shed light on the interplay between radiation, dust dynamics, and star formation, offering valuable insights into the complex processes that shape stars, their environments, and compositions within molecular clouds.

Support for for NS and PFH was provided by NSF Research Grants 1911233, 20009234, 2108318, NSF CAREER grant 1455342, NASA grants 80NSSC18K0562, HST-AR-15800. Support for MYG was provided by NASA through the NASA Hubble Fellowship grant #HST-HF2-51479 awarded by the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., for NASA, under contract NAS5-26555. Numerical calculations were run on the TACC compute cluster “Frontera,” allocations AST21010, AST20016, and AST21002 supported by the NSF and TACC, and NASA HEC SMD-16-7592. This research is part of the Frontera computing project at the Texas Advanced Computing Center. Frontera is made possible by National Science Foundation award OAC-1818253.

Data Availability Statement

The data supporting this article are available on reasonable request to the corresponding author.

References

  • Abergel et al. (2002) Abergel, A., Bernard, J. P., Boulanger, F., et al. 2002, A&A, 389, 239, doi: 10.1051/0004-6361:20020324
  • Ali (2021) Ali, A. A. 2021, Monthly Notices of the Royal Astronomical Society, 501, 4136
  • Altobelli et al. (2007) Altobelli, N., Dikarev, V., Kempf, S., et al. 2007, Journal of Geophysical Research (Space Physics), 112, 7105, doi: 10.1029/2006JA011978
  • Altobelli et al. (2006) Altobelli, N., Grün, E., & Landgraf, M. 2006, A&A, 448, 243, doi: 10.1051/0004-6361:20053909
  • Bai & Stone (2010) Bai, X.-N., & Stone, J. M. 2010, ApJ, 722, L220, doi: 10.1088/2041-8205/722/2/L220
  • Biazzo et al. (2015) Biazzo, K., Gratton, R., Desidera, S., et al. 2015, Astronomy & Astrophysics, 583, A135
  • Boogert et al. (2013) Boogert, A., Chiar, J., Knez, C., et al. 2013, The Astrophysical Journal, 777, 73
  • Carballido et al. (2008) Carballido, A., Stone, J. M., & Turner, N. J. 2008, MNRAS, 386, 145, doi: 10.1111/j.1365-2966.2008.13014.x
  • Cochran & Ostriker (1977) Cochran, W. D., & Ostriker, J. P. 1977, Astrophysical Journal, vol. 211, Jan. 15, 1977, pt. 1, p. 392-399. Research supported by the Aspen Center for Physics, 211, 392
  • Dorschner (2003) Dorschner, J. 2003, in Lecture Notes in Physics, Berlin Springer Verlag, Vol. 609, Astromineralogy; University Observatory Schillergässchen 3, D-07745 Jena, Germany, ed. T. K. Henning, 1–54
  • Draine & Lee (1984) Draine, B., & Lee, H. M. 1984, Astrophysical Journal, Part 1 (ISSN 0004-637X), vol. 285, Oct. 1, 1984, p. 89-108. Research supported by the Alfred P. Sloan Foundation., 285, 89
  • Draine (2003) Draine, B. T. 2003, ARA&A, 41, 241, doi: 10.1146/annurev.astro.41.011802.094840
  • Draine & Sutin (1987) Draine, B. T., & Sutin, B. 1987, ApJ, 320, 803, doi: 10.1086/165596
  • Flagey et al. (2009) Flagey, N., Noriega-Crespo, A., Boulanger, F., et al. 2009, The Astrophysical Journal, 701, 1450
  • Frisch & Slavin (2003) Frisch, P. C., & Slavin, J. D. 2003, ApJ, 594, 844, doi: 10.1086/376689
  • Girichidis et al. (2020) Girichidis, P., Offner, S. S., Kritsuk, A. G., et al. 2020, Space Science Reviews, 216, 1
  • Grudić et al. (2022) Grudić, M. Y., Guszejnov, D., Offner, S. S. R., et al. 2022, MNRAS, 512, 216, doi: 10.1093/mnras/stac526
  • Grudić et al. (2021) Grudić, M. Y., Guszejnov, D., Hopkins, P. F., Offner, S. S. R., & Faucher-Giguère, C.-A. 2021, Monthly Notices of the Royal Astronomical Society, 506, 2199, doi: 10.1093/mnras/stab1347
  • Gustafsson (2018a) Gustafsson, B. 2018a, Astronomy & Astrophysics, 616, A91
  • Gustafsson (2018b) —. 2018b, Astronomy & Astrophysics, 620, A53
  • Höfner & Olofsson (2018) Höfner, S., & Olofsson, H. 2018, A&A Rev., 26, 1, doi: 10.1007/s00159-017-0106-5
  • Hopkins (2014) Hopkins, P. F. 2014, ApJ, 797, 59, doi: 10.1088/0004-637X/797/1/59
  • Hopkins (2015) —. 2015, MNRAS, 450, 53, doi: 10.1093/mnras/stv195
  • Hopkins (2016) —. 2016, MNRAS, 462, 576, doi: 10.1093/mnras/stw1578
  • Hopkins & Conroy (2017) Hopkins, P. F., & Conroy, C. 2017, ApJ, 835, 154, doi: 10.3847/1538-4357/835/2/154
  • Hopkins & Grudić (2019a) Hopkins, P. F., & Grudić, M. Y. 2019a, MNRAS, 483, 4187, doi: 10.1093/mnras/sty3089
  • Hopkins & Grudić (2019b) —. 2019b, MNRAS, 483, 4187, doi: 10.1093/mnras/sty3089
  • Hopkins et al. (2020) Hopkins, P. F., Grudić, M. Y., Wetzel, A., et al. 2020, MNRAS, 491, 3702, doi: 10.1093/mnras/stz3129
  • Hopkins & Lee (2016) Hopkins, P. F., & Lee, H. 2016, MNRAS, 456, 4174, doi: 10.1093/mnras/stv2745
  • Hopkins & Raives (2016) Hopkins, P. F., & Raives, M. J. 2016, MNRAS, 455, 51, doi: 10.1093/mnras/stv2180
  • Hopkins et al. (2022) Hopkins, P. F., Rosen, A. L., Squire, J., et al. 2022, Monthly Notices of the Royal Astronomical Society, 517, 1491
  • Hopkins et al. (2023) Hopkins, P. F., Wetzel, A., Wheeler, C., et al. 2023, MNRAS, 519, 3154, doi: 10.1093/mnras/stac3489
  • Johansen et al. (2009) Johansen, A., Youdin, A., & Mac Low, M.-M. 2009, ApJ, 704, L75, doi: 10.1088/0004-637X/704/2/L75
  • King & Pounds (2015) King, A., & Pounds, K. 2015, Annual Review of Astronomy and Astrophysics, 53, 115
  • Krause et al. (2020) Krause, M. G. H., Offner, S. S. R., Charbonnel, C., et al. 2020, Space Sci. Rev., 216, 64, doi: 10.1007/s11214-020-00689-4
  • Krüger et al. (2001) Krüger, H., et al. 2001, Planetary and Space Science, 49, 1303, doi: 10.1016/S0032-0633(01)00054-X
  • Krumholz (2018) Krumholz, M. R. 2018, MNRAS, 480, 3468, doi: 10.1093/mnras/sty2105
  • Krumholz et al. (2009) Krumholz, M. R., Klein, R. I., McKee, C. F., Offner, S. S., & Cunningham, A. J. 2009, Science, 323, 754
  • Krumholz & Thompson (2012) Krumholz, M. R., & Thompson, T. A. 2012, ApJ, 760, 155, doi: 10.1088/0004-637X/760/2/155
  • Larson (1981) Larson, R. B. 1981, Monthly Notices of the Royal Astronomical Society, 194, 809
  • Lee et al. (2017) Lee, H., Hopkins, P. F., & Squire, J. 2017, MNRAS, 469, 3532, doi: 10.1093/mnras/stx1097
  • Li & Draine (2001) Li, A., & Draine, B. 2001, The Astrophysical Journal, 554, 778
  • Lupi et al. (2018) Lupi, A., Bovino, S., Capelo, P. R., Volonteri, M., & Silk, J. 2018, MNRAS, 474, 2884, doi: 10.1093/mnras/stx2874
  • Lupi et al. (2017) Lupi, A., Volonteri, M., & Silk, J. 2017, MNRAS, 470, 1673, doi: 10.1093/mnras/stx1313
  • Mac Low & Klessen (2004) Mac Low, M.-M., & Klessen, R. S. 2004, Reviews of modern physics, 76, 125
  • Maia et al. (2014) Maia, M. T., Meléndez, J., & Ramírez, I. 2014, The Astrophysical Journal Letters, 790, L25
  • Mathews (1967) Mathews, W. G. 1967, The Astrophysical Journal, 147, 965
  • Mathis (1990) Mathis, J. S. 1990, Annual Review of Astronomy and Astrophysics, 28, 37
  • Mathis et al. (1977) Mathis, J. S., Rumpl, W., & Nordsieck, K. H. 1977, ApJ, 217, 425, doi: 10.1086/155591
  • McKee & Offner (2010) McKee, C. F., & Offner, S. S. R. 2010, ApJ, 716, 167, doi: 10.1088/0004-637X/716/1/167
  • McKee & Ostriker (2007) McKee, C. F., & Ostriker, E. C. 2007, Annu. Rev. Astron. Astrophys., 45, 565
  • McKinnon et al. (2018) McKinnon, R., Vogelsberger, M., Torrey, P., Marinacci, F., & Kannan, R. 2018, MNRAS, 478, 2851, doi: 10.1093/mnras/sty1248
  • Meisel et al. (2002) Meisel, D. D., Janches, D., & Mathews, J. D. 2002, ApJ, 579, 895, doi: 10.1086/342919
  • Melendez et al. (2009) Melendez, J., Asplund, M., Gustafsson, B., & Yong, D. 2009, The Astrophysical Journal, 704, L66
  • Minissale et al. (2016) Minissale, M., Dulieu, F., Cazaux, S., & Hocuk, S. 2016, Astronomy & Astrophysics, 585, A24
  • Moseley et al. (2019) Moseley, E. R., Squire, J., & Hopkins, P. F. 2019, MNRAS, 489, 325, doi: 10.1093/mnras/stz2128
  • Murray et al. (2005) Murray, N., Quataert, E., & Thompson, T. A. 2005, The Astrophysical Journal, 618, 569
  • Nissen et al. (2017) Nissen, P., Aguirre, V. S., Christensen-Dalsgaard, J., et al. 2017, Astronomy & Astrophysics, 608, A112
  • Oh et al. (2018) Oh, S., Price-Whelan, A. M., Brewer, J. M., et al. 2018, The Astrophysical Journal, 854, 138
  • Pan et al. (2011) Pan, L., Padoan, P., Scalo, J., Kritsuk, A. G., & Norman, M. L. 2011, ApJ, 740, 6, doi: 10.1088/0004-637X/740/1/6
  • Poppe et al. (2010) Poppe, A., James, D., Jacobsmeyer, B., & Horányi, M. 2010, Geophysical Research Letters, 37, 11101, doi: 10.1029/2010GL043300
  • Ramírez et al. (2015) Ramírez, I., Khanal, S., Aleo, P., et al. 2015, The Astrophysical Journal, 808, 13
  • Rosen et al. (2016) Rosen, A. L., Krumholz, M. R., McKee, C. F., & Klein, R. I. 2016, Monthly Notices of the Royal Astronomical Society, 463, 2553
  • Saffe et al. (2017) Saffe, C., Jofré, E., Martioli, E., et al. 2017, Astronomy & Astrophysics, 604, L4
  • Salpeter (1977) Salpeter, E. 1977, Annual Review of Astronomy and Astrophysics, 15, 267
  • Shu (1977) Shu, F. H. 1977, ApJ, 214, 488, doi: 10.1086/155274
  • Soliman & Hopkins (2023) Soliman, N. H., & Hopkins, P. F. 2023, Monthly Notices of the Royal Astronomical Society, 525, 2668
  • Spitzer Jr (2008) Spitzer Jr, L. 2008, Physical processes in the interstellar medium (John Wiley & Sons)
  • Teske et al. (2016a) Teske, J. K., Khanal, S., & Ramírez, I. 2016a, The Astrophysical Journal, 819, 19
  • Teske et al. (2016b) Teske, J. K., Shectman, S. A., Vogt, S. S., et al. 2016b, The Astronomical Journal, 152, 167
  • Thoraval et al. (1997) Thoraval, S., Boisse, P., & Duvert, G. 1997, A&A, 319, 948
  • Thoraval et al. (1999) Thoraval, S., Boissé, P., & Duvert, G. 1999, A&A, 351, 1051
  • Tielens (2005) Tielens, A. G. G. M. 2005, The Physics and Chemistry of the Interstellar Medium (Cambridge, UK: Cambridge University Press)
  • Toci et al. (2021) Toci, C., Rosotti, G., Lodato, G., Testi, L., & Trapman, L. 2021, Monthly Notices of the Royal Astronomical Society, 507, 818
  • Tout et al. (1996) Tout, C. A., Pols, O. R., Eggleton, P. P., & Han, Z. 1996, MNRAS, 281, 257, doi: 10.1093/mnras/281.1.257
  • Watanabe & Kouchi (2008) Watanabe, N., & Kouchi, A. 2008, Progress In Surface Science, 83, 439
  • Weingartner & Draine (2001) Weingartner, J. C., & Draine, B. 2001, The Astrophysical Journal, 563, 842
  • Weingartner & Draine (2001) Weingartner, J. C., & Draine, B. T. 2001, ApJS, 134, 263, doi: 10.1086/320852
  • Whittet et al. (1993) Whittet, D., Millar, T., & Williams, D. 1993, Dust and Chemistry in Astronomy, Institute of Physics Publishing, Bristol and Philadelphia