License: arXiv.org perpetual non-exclusive license
arXiv:2403.03437v2 [astro-ph.GA] 18 Mar 2024

Dark Dragon Breaks Magnetic Chain: Dynamical Substructures of IRDC G28.34 Form in Supported Environments

Junhao Liu (刘峻豪) National Astronomical Observatory of Japan, 2-21-1 Osawa, Mitaka, Tokyo 181-8588, Japan East Asian Observatory, 660 N. A‘ohōkū Place, University Park, Hilo, HI 96720, USA Qizhou Zhang Center for Astrophysics |||| Harvard & Smithsonian, 60 Garden Street, Cambridge, MA 02138, USA Yuxin Lin Max-Planck-Institut für Extraterrestrische Physik, Giessenbachstr 1, D-85748 Garching bei München, Germany Ke** Qiu School of Astronomy and Space Science, Nan**g University, 163 Xianlin Avenue, Nan**g 210023, Jiangsu, People’s Republic of China Key Laboratory of Modern Astronomy and Astrophysics (Nan**g University), Ministry of Education, Nan**g 210023, Jiangsu, People’s Republic of China Patrick M. Koch Institute of Astronomy and Astrophysics, Academia Sinica, 11F of Astronomy-Mathematics Building, AS/NTU No.1, Sec. 4, Roosevelt Rd, Taipei 10617, Taiwan, Republic of China Hauyu Baobab Liu Department of Physics, National Sun Yat-Sen University, No. 70, Lien-Hai Road, Kaohsiung City 80424, Taiwan, Republic of China Center of Astronomy and Gravitation, National Taiwan Normal University, Taipei 116, Taiwan, Republic of China Zhi-Yun Li Astronomy Department, University of Virginia, Charlottesville, VA 22904-4325, USA Josep Miquel Girart Institut de Ciències de l’Espai (ICE, CSIC), Can Magrans s/n, E-08193 Cerdanyola del Vallès, Catalonia, Spain Institut d’Estudis Espacials de de Catalunya (IEEC), E-08034 Barcelona, Catalonia, Spain Thushara G.S. Pillai MIT Haystack Observatory, 99 Millstone Road, Westford, MA, 01826, USA Shanghuo Li Max Planck Institute for Astronomy, Konigstuhl 17, D-69117 Heidelberg, Germany Huei-Ru Vivien Chen Institute of Astronomy and Astrophysics, Academia Sinica, 11F of Astronomy-Mathematics Building, AS/NTU No.1, Sec. 4, Roosevelt Rd, Taipei 10617, Taiwan, Republic of China Tao-Chung Ching National Radio Astronomy Observatory, P.O. Box O, Socorro, NM 87801, USA Paul T. P. Ho East Asian Observatory, 660 N. A‘ohōkū Place, University Park, Hilo, HI 96720, USA Institute of Astronomy and Astrophysics, Academia Sinica, 11F of Astronomy-Mathematics Building, AS/NTU No.1, Sec. 4, Roosevelt Rd, Taipei 10617, Taiwan, Republic of China Shih-** Lai Institute of Astronomy and Department of Physics, National Tsing Hua University, Hsinchu 30013, Taiwan, Republic of China Institute of Astronomy and Astrophysics, Academia Sinica, 11F of Astronomy-Mathematics Building, AS/NTU No.1, Sec. 4, Roosevelt Rd, Taipei 10617, Taiwan, Republic of China Ramprasad Rao Center for Astrophysics |||| Harvard & Smithsonian, 60 Garden Street, Cambridge, MA 02138, USA Institute of Astronomy and Astrophysics, Academia Sinica, 11F of Astronomy-Mathematics Building, AS/NTU No.1, Sec. 4, Roosevelt Rd, Taipei 10617, Taiwan, Republic of China Ya-Wen Tang Institute of Astronomy and Astrophysics, Academia Sinica, 11F of Astronomy-Mathematics Building, AS/NTU No.1, Sec. 4, Roosevelt Rd, Taipei 10617, Taiwan, Republic of China Ke Wang Kavli Institute for Astronomy and Astrophysics, Peking University, 5 Yiheyuan Road, Haidian District, Bei**g 100871, People’s Republic of China
Abstract

We have comprehensively studied the multi-scale physical properties of the infrared dark cloud (IRDC) G28.34 (the Dragon cloud) with dust polarization and molecular line data from Planck, FCRAO-14m, JCMT, and ALMA. We find that the averaged magnetic fields of clumps tend to be either parallel with or perpendicular to the cloud-scale magnetic fields, while the cores in clump MM4 tend to have magnetic fields aligned with the clump fields. Implementing the relative orientation analysis (for magnetic fields, column density gradients, and local gravity), Velocity Gradient Technique (VGT), and modified Davis-Chandrasekhar-Fermi (DCF) analysis, we find that: G28.34 is located in a trans-to-sub-Alfvénic environment (A=0.74subscript𝐴0.74\mathcal{M}_{A}=0.74caligraphic_M start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT = 0.74 within r=15𝑟15r=15italic_r = 15 pc); the magnetic field is effectively resisting gravitational collapse in large-scale diffuse gas, but is distorted by gravity within the cloud and affected by star formation activities in high-density regions; and the normalized mass-to-flux ratio tends to increase with increasing density and decreasing radius. Considering the thermal, turbulent, and magnetic supports, we find that the environmental gas of G28.34 is in a super-virial (supported) state, the infrared dark clumps may be in a near-equilibrium state, and core MM4-core4 is in a sub-virial (gravity-dominant) state. In summary, we suggest that magnetic fields dominate gravity and turbulence in the cloud environment at large scales, resulting in relatively slow cloud formation and evolution processes. Within the cloud, gravity could overwhelm magnetic fields and turbulence, allowing local dynamical star formation to happen.

Star formation (1569) — Molecular clouds (1072) — Interstellar medium (847) — Magnetic fields (994)
facilities: Planck, JCMT, ALMAsoftware: Astropy (Astropy Collaboration et al., 2013, 2018), Matplotlib (Hunter, 2007).

1 Introduction

Magnetic fields and turbulence are two major forces resisting the gravitational collapse of molecular clouds in star formation regions (McKee & Ostriker, 2007). Observational studies of the magnetic field are important to understand how it regulates star formation and how it is affected by star formation. Observations of polarized dust emission (produced by dust grain alignment, Lazarian, 2007; Lazarian & Hoang, 2007; Andersson et al., 2015) have been the most widely used technique to trace the plane-of-sky (POS) magnetic field orientation in star-forming molecular clouds (e.g., Hildebrand et al., 1984). There has been an increasing number of both single-dish and interferometric dust polarization observations in molecular clouds111Cloud, clump, core, and condensation scales corresponds to similar-to\sim10 pc, similar-to\sim1 pc, similar-to\sim0.1 pc, and similar-to\sim0.01 pc, respectively. (Pattle & Fissel, 2019; Hull & Zhang, 2019). From the observational magnetic field studies, the recent review papers summarised that magnetically trans-to-super-critical clumps/cores form in sub-critical and trans-to-sub-Alfvénic clouds (Liu et al., 2022a, b). The substructures of clouds may transit to an averagely trans-to-super-Alfvénic state as density increases (Li, 2021; Liu et al., 2022b), but this result is less clear due to the uncertainties of the analysis methods. Despite the progress, most of the previous observational studies of magnetic fields were in more evolved star formation regions where significant star-forming activities have taken place (e.g., Sanhueza et al., 2021; Liu et al., 2023a). The general magnetic field properties of clouds at early star formation stages remain under explored.

The massive star formation process is relatively less understood than low-mass star formation partly due to a lack of observations at early evolutionary stages. Massive infrared dark clouds (IRDCs) are believed to harbor the early phase of massive star formation (Pillai et al., 2006; Rathborne et al., 2006; Sanhueza et al., 2012, 2019), but their weak polarized dust emission makes it more challenging to detect than more evolved regions. So far, there have been only a handful of single-dish (Pillai et al., 2015; Juvela et al., 2018; Liu et al., 2018; Tang et al., 2019; Soam et al., 2019; Añez-López et al., 2020; Ngoc et al., 2023) and interferometric (Beuther et al., 2018; Cortes et al., 2019; Liu et al., 2020) studies of magnetic fields in IRDCs, and there is a lack of multi-scale studies of magnetic fields in the same IRDC. Since the dynamic role of the magnetic field may vary from large to small scales (Liu et al., 2022a, b), it is essential to comprehensively investigate the multi-scale magnetic field in a single IRDC to advance our understanding of the magnetic field properties in the early stages of massive star formation.

G28.34+0.06 (hereafter G28.34, also known as the Dragon cloud) is a well-studied massive filamentary IRDC located at a distance of 4.8 kpc (e.g., Carey et al., 1998; Pillai et al., 2006; Rathborne et al., 2006; Wang et al., 2008; Zhang et al., 2009; Lin et al., 2017; Wang, 2018). The majority of G28.34 is 8 μ𝜇\muitalic_μm-dark except for the northern end (Zhang et al., 2009). Rathborne et al. (2006) has identified 18 millimeter dust continuum clumps (MM1-18) within G28.34 and in its vicinity. The clumps in G28.34 are found to further fragment into smaller substructures with higher resolution observations (e.g., Zhang et al., 2009, 2015; Kong, 2019). All the clumps in G28.34 show signs of star formation activities (e.g., Wang et al., 2006; Kong et al., 2019). The large mass reservoir and infrared dark behavior make G28.34 a perfect place to study the extremely early evolutionary stage of massive star formation.

Liu et al. (2020) presented a study of the small-scale magnetic field in three clumps (MM1, MM4, and MM9) in G28.34 with Atacama Large Millimeter/submillimeter Array (ALMA) polarization observations. They found that even considering both the turbulent and magnetic support, core MM4-core4 is still in a non-equilibrium state dominated by gravity. As a follow-up work of Liu et al. (2020), in this paper, we utilize the James Clerk Maxwell Telescope (JCMT) dust polarization observations and the Planck dust polarization data to study the multi-scale magnetic field in G28.34 and to determine the multi-scale energy balance in this IRDC.

2 Observations

2.1 ALMA dust polarization and molecular line observations

Clumps MM4 and MM9 in G28.34 were observed by ALMA between 2017 April 18 and 2018 September 11 under projects 2016.1.00248.S (PI: Qizhou Zhang) and 2017.1.00793.S (PI: Qizhou Zhang) in configurations C-1, C-3, and C-4. Three spectral windows were configured to observe the dust continuum at similar-to\sim215.5–219.5 GHz and similar-to\sim232.5–234.5 GHz (band 6) in the full polarization mode. Four spectral windows were configured to cover the 1212{}^{12}start_FLOATSUPERSCRIPT 12 end_FLOATSUPERSCRIPTCO (2-1), OCS (19-18), 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCS (5-4), and N22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPTD+{}^{+}start_FLOATSUPERSCRIPT + end_FLOATSUPERSCRIPT (3-2) lines with a channel width of 122 kHz (0.16 km s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT) over a bandwidth of 58.6 MHz (similar-to\sim76 km s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT). The data in configurations C-1 and C-3 have been previously reported in Liu et al. (2020).

The data were calibrated with Common Astronomy Software Applications (CASA, McMullin et al., 2007). We performed two rounds of phase-only self-calibrations for the dust continuum. We imaged the molecular line cubes and Stokes I𝐼Iitalic_I, Q𝑄Qitalic_Q, and U𝑈Uitalic_U maps of dust continuum using the CASA task TCLEAN with a Briggs weighting parameter of robust = 0.5. We adopted a pixel size of 0.10.10.1\arcsec0.1 ″ for the imaging. The synthesized beam of the three configurations-combined images is 0.7×0.5similar-toabsent0.70.5\sim 0.7\arcsec\times 0.5\arcsec∼ 0.7 ″ × 0.5 ″ (similar-to\sim0.016-0.012 pc at a distance of 4.8 kpc), which improves the resolution of our previous two configurations-combined images (0.9×0.7similar-toabsent0.90.7\sim 0.9\arcsec\times 0.7\arcsec∼ 0.9 ″ × 0.7 ″, Liu et al., 2020). The maximum recoverable scale222https://almascience.eso.org/observing/observing-configuration-schedule/prior-cycle-observing-and-configuration-schedule is 13similar-toabsent13\sim 13\arcsec∼ 13 ″ (similar-to\sim0.3 pc at 4.8 kpc). Before primary beam correction, the 1σ𝜎\sigmaitalic_σ root-mean-square (RMS) noise is σIsimilar-tosubscript𝜎𝐼absent\sigma_{I}\simitalic_σ start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT ∼0.03 and 0.03 mJy beam11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT for the Stokes I𝐼Iitalic_I dust continuum maps and σQUsimilar-tosubscript𝜎𝑄𝑈absent\sigma_{QU}\simitalic_σ start_POSTSUBSCRIPT italic_Q italic_U end_POSTSUBSCRIPT ∼0.01 and 0.012 mJy beam11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT for the Stokes Q𝑄Qitalic_Q or U𝑈Uitalic_U dust continuum maps of MM4 and MM9, respectively. The debiased polarized intensity PI𝑃𝐼PIitalic_P italic_I and its corresponding uncertainty σPIsubscript𝜎𝑃𝐼\sigma_{PI}italic_σ start_POSTSUBSCRIPT italic_P italic_I end_POSTSUBSCRIPT are calculated as PI=Q2+U2σQU2𝑃𝐼superscript𝑄2superscript𝑈2superscriptsubscript𝜎𝑄𝑈2PI=\sqrt{Q^{2}+U^{2}-\sigma_{QU}^{2}}italic_P italic_I = square-root start_ARG italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_U start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_σ start_POSTSUBSCRIPT italic_Q italic_U end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG (Vaillancourt, 2006) and σPI2σQUsimilar-tosubscript𝜎𝑃𝐼2subscript𝜎𝑄𝑈\sigma_{PI}\sim\sqrt{2}\sigma_{QU}italic_σ start_POSTSUBSCRIPT italic_P italic_I end_POSTSUBSCRIPT ∼ square-root start_ARG 2 end_ARG italic_σ start_POSTSUBSCRIPT italic_Q italic_U end_POSTSUBSCRIPT, where σQUsubscript𝜎𝑄𝑈\sigma_{QU}italic_σ start_POSTSUBSCRIPT italic_Q italic_U end_POSTSUBSCRIPT is the 1σ𝜎\sigmaitalic_σ rms noise on the background region of the Q𝑄Qitalic_Q or U𝑈Uitalic_U maps. The polarization position angle θpsubscript𝜃p\theta_{\mathrm{p}}italic_θ start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT is estimated with θp=0.5arctan(U/Q)subscript𝜃p0.5𝑈𝑄\theta_{\mathrm{p}}=0.5\arctan(U/Q)italic_θ start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT = 0.5 roman_arctan ( italic_U / italic_Q ). The uncertainty on the polarization position angle (Naghizadeh-Khouei & Clarke, 1993) is given by δθ=0.5σQU2/(Q2+U2)20°.26(σPI/PI)28°.65(σQU/PI)\delta\theta=0.5\sqrt{\sigma_{QU}^{2}/(Q^{2}+U^{2}})\sim 20\arcdeg.26(\sigma_{% PI}/PI)\sim 28\arcdeg.65(\sigma_{QU}/PI)italic_δ italic_θ = 0.5 square-root start_ARG italic_σ start_POSTSUBSCRIPT italic_Q italic_U end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ( italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_U start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) ∼ 20 ° .26 ( italic_σ start_POSTSUBSCRIPT italic_P italic_I end_POSTSUBSCRIPT / italic_P italic_I ) ∼ 28 ° .65 ( italic_σ start_POSTSUBSCRIPT italic_Q italic_U end_POSTSUBSCRIPT / italic_P italic_I ). We only adopt the N22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPTD+{}^{+}start_FLOATSUPERSCRIPT + end_FLOATSUPERSCRIPT (3-2) line data in this study. The RMS noises of the N22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPTD+{}^{+}start_FLOATSUPERSCRIPT + end_FLOATSUPERSCRIPT line cubes (before primary beam correction) are similar-to\sim1.6 and 1.9 mJy beam11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT at a 0.16 km s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT channel for MM4 and MM9, respectively. All the ALMA images shown in this paper are before primary beam correction. All the continuum fluxes used for calculations of physical parameters are after primary beam correction.

2.2 JCMT dust polarization observations and molecular line data

The 850 μ𝜇\muitalic_μm polarized emission of G28.34 was observed by SCUBA-2/POL-2 (Holland et al., 2013; Friberg et al., 2016) on the JCMT between 2022 February 24 and 2022 June 25 under the project M22AP018 (PI: Junhao Liu). The observations were made with the POL-2 DAISY mode with low noise levels in a central region of a 3\arcmin radius and with increased noises toward the edge. The spatial resolution of JCMT is 14similar-toabsent14\sim 14\arcsec∼ 14 ″ (similar-to\sim0.33 pc) at 850 μ𝜇\muitalic_μm. The center of our DAISY field is in the southern part of G28.34 near MM4. A similar POL-2 observation centered at MM1 has been conducted by the JCMT large program B-Fields in STar-Forming Region Observations (BISTRO). A detailed analysis of the multi-scale magnetic field in the infrared bright clump MM1 will be presented by Hwang, J. et al. in prep. as part of the BISTRO survey.

The POL-2 data were reduced using the SMURF (Jenness et al., 2013) package of Starlink (Currie et al., 2014) in a process similar to Liu et al. (2019). The final maps were gridded to 7\arcsec pixels (Nyquist sampling). The flux conversion factor (FCF) is 495 Jy beam11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT pW11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT for SCUBA-2 after 2018 June 30 (Mairs et al., 2021). Accounting for the additional flux losses of the POL-2 with a factor of 1.35, we adopt an FCF of 668 Jy beam11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT pW11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT to convert the unit of our POL-2 data from pW to Jy beam11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT. Within the central 3\arcmin-radius region, the mean values for the observational uncertainty of I𝐼Iitalic_I, Q𝑄Qitalic_Q, and U𝑈Uitalic_U (i.e., δI𝛿𝐼\delta Iitalic_δ italic_I, δQ𝛿𝑄\delta Qitalic_δ italic_Q, and δU𝛿𝑈\delta Uitalic_δ italic_U) are similar-to\sim1.8, 1.9, and 1.9 mJy beam11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT, respectively. The debiased polarized intensity PI𝑃𝐼PIitalic_P italic_I and its corresponding error δPI𝛿𝑃𝐼\delta PIitalic_δ italic_P italic_I are calculated as PI=Q2+U20.5(δQ2+δU2)𝑃𝐼superscript𝑄2superscript𝑈20.5𝛿superscript𝑄2𝛿superscript𝑈2PI=\sqrt{Q^{2}+U^{2}-0.5(\delta Q^{2}+\delta U^{2})}italic_P italic_I = square-root start_ARG italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_U start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 0.5 ( italic_δ italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_δ italic_U start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_ARG and δPI=(QδQ+UδU)/Q2+U2𝛿𝑃𝐼𝑄𝛿𝑄𝑈𝛿𝑈superscript𝑄2superscript𝑈2\delta PI=(Q\delta Q+U\delta U)/\sqrt{Q^{2}+U^{2}}italic_δ italic_P italic_I = ( italic_Q italic_δ italic_Q + italic_U italic_δ italic_U ) / square-root start_ARG italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_U start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG, respectively. The polarization position angle θpsubscript𝜃p\theta_{\mathrm{p}}italic_θ start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT and its uncertainty δθ𝛿𝜃\delta\thetaitalic_δ italic_θ (Naghizadeh-Khouei & Clarke, 1993) are estimated with θp=0.5arctan(U/Q)subscript𝜃p0.5𝑈𝑄\theta_{\mathrm{p}}=0.5\arctan(U/Q)italic_θ start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT = 0.5 roman_arctan ( italic_U / italic_Q ) and δθ=0.5(Q2δU2+U2δQ2)/(Q2+U2)228°.65(δPI/PI)𝛿𝜃0.5superscript𝑄2𝛿superscript𝑈2superscript𝑈2𝛿superscript𝑄2superscriptsuperscript𝑄2superscript𝑈22similar-to28°.65𝛿𝑃𝐼𝑃𝐼\delta\theta=0.5\sqrt{(Q^{2}\delta U^{2}+U^{2}\delta Q^{2})/(Q^{2}+U^{2})^{2}}% \sim 28\arcdeg.65(\delta PI/PI)italic_δ italic_θ = 0.5 square-root start_ARG ( italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_δ italic_U start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_U start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_δ italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) / ( italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_U start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ∼ 28 ° .65 ( italic_δ italic_P italic_I / italic_P italic_I ), respectively, where δPIδQδUsimilar-to𝛿𝑃𝐼𝛿𝑄similar-to𝛿𝑈\delta PI\sim\delta Q\sim\delta Uitalic_δ italic_P italic_I ∼ italic_δ italic_Q ∼ italic_δ italic_U. The polarization percentage is given by P=PI/I𝑃𝑃𝐼𝐼P=PI/Iitalic_P = italic_P italic_I / italic_I and δP=δPI2/I2+δI2(Q2+U2)/I4𝛿𝑃𝛿𝑃superscript𝐼2superscript𝐼2𝛿superscript𝐼2superscript𝑄2superscript𝑈2superscript𝐼4\delta P=\sqrt{\delta PI^{2}/I^{2}+\delta I^{2}(Q^{2}+U^{2})/I^{4}}italic_δ italic_P = square-root start_ARG italic_δ italic_P italic_I start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_I start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_δ italic_I start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_U start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) / italic_I start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT end_ARG (see Appendix A).

Additionally, we collect the archival 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO (3-2) (program: M10AC06) and HCO+{}^{+}start_FLOATSUPERSCRIPT + end_FLOATSUPERSCRIPT (4-3) (programs: M16BP081 and M17BP087) line data observed with the Heterodyne Array Receiver Program (HARP) and Auto-Correlation Spectrometer and Imaging System (ACSIS, Buckle et al., 2009) on the JCMT. The spatial and spectral resolutions are 14similar-toabsent14\sim 14\arcsec∼ 14 ″ and 0.055 km s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT, respectively, for both lines. To increase the signal-to-noise ratio (S/N), we smooth the two lines to a spectral resolution of 0.2 km s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT. We estimate the main beam temperature (Tmbsubscript𝑇mbT_{\mathrm{mb}}italic_T start_POSTSUBSCRIPT roman_mb end_POSTSUBSCRIPT) from the corrected antenna temperature (TAsubscriptsuperscript𝑇AT^{\ast}_{\mathrm{A}}italic_T start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_A end_POSTSUBSCRIPT) adopting a main beam efficiency of 0.64 (Buckle et al., 2009). The typical observation uncertainties of 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO (3-2) and HCO+{}^{+}start_FLOATSUPERSCRIPT + end_FLOATSUPERSCRIPT (4-3) are 0.5-1 K and 0.2-0.4 K (in Tmbsubscript𝑇mbT_{\mathrm{mb}}italic_T start_POSTSUBSCRIPT roman_mb end_POSTSUBSCRIPT), respectively, per 0.2 km s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT channel within our studied region.

2.3 Planck 353 GHz dust polarization data

We adopt the Planck High Frequency Instrument (HFI, Lamarre et al., 2010) 353 GHz Stokes I𝐼Iitalic_I, Q𝑄Qitalic_Q, and U𝑈Uitalic_U maps of the thermal dust emission (version R3.00, Planck Collaboration et al., 2020) toward G28.34 and its surrounding area constructed with the Generalized Needlet Internal Linear Combination method (GNILC, Remazeilles et al., 2011). We also adopt the earlier released dust optical depth (τ353subscript𝜏353\tau_{353}italic_τ start_POSTSUBSCRIPT 353 end_POSTSUBSCRIPT) and temperature maps (version R1.02, Planck Collaboration et al., 2014). The resolution of the Planck maps is 5\arcmin (similar-to\sim7 pc) at 353 GHz. The pixel size of the Planck maps is 1.71\arcmin. Within our considered map area (1°×1°1°1°1\arcdeg\times 1\arcdeg1 ° × 1 °), the mean values for the uncertainties of Q𝑄Qitalic_Q and U𝑈Uitalic_U (i.e., δQ𝛿𝑄\delta Qitalic_δ italic_Q and δU𝛿𝑈\delta Uitalic_δ italic_U) are similar-to\sim3 and 4 μ𝜇\muitalic_μKCMBCMB{}_{\mathrm{CMB}}start_FLOATSUBSCRIPT roman_CMB end_FLOATSUBSCRIPT, respectively. The debiased polarized intensity PI𝑃𝐼PIitalic_P italic_I and its corresponding uncertainty δPI𝛿𝑃𝐼\delta PIitalic_δ italic_P italic_I are calculated as PI=Q2+U20.5(δQ2+δU2)𝑃𝐼superscript𝑄2superscript𝑈20.5𝛿superscript𝑄2𝛿superscript𝑈2PI=\sqrt{Q^{2}+U^{2}-0.5(\delta Q^{2}+\delta U^{2})}italic_P italic_I = square-root start_ARG italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_U start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 0.5 ( italic_δ italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_δ italic_U start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_ARG and δPI(Q2δQ2+U2δU2)/(Q2+U2)similar-to𝛿𝑃𝐼superscript𝑄2𝛿superscript𝑄2superscript𝑈2𝛿superscript𝑈2superscript𝑄2superscript𝑈2\delta PI\sim\sqrt{(Q^{2}\delta Q^{2}+U^{2}\delta U^{2})/(Q^{2}+U^{2})}italic_δ italic_P italic_I ∼ square-root start_ARG ( italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_δ italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_U start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_δ italic_U start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) / ( italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_U start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_ARG, respectively. We compute polarization position angles in the equatorial coordinates with θp=0.5arctan(U/Q)Δθpgesubscript𝜃p0.5𝑈𝑄Δsuperscriptsubscript𝜃pge\theta_{\mathrm{p}}=0.5\arctan(U/Q)-\Delta\theta_{\mathrm{p}}^{\mathrm{g-e}}italic_θ start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT = 0.5 roman_arctan ( italic_U / italic_Q ) - roman_Δ italic_θ start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_g - roman_e end_POSTSUPERSCRIPT, where

Δθpge=arctan(cos(l32.9°)cosbcot62.9°sinbsin(l32.9°))Δsuperscriptsubscript𝜃pge𝑙32.9°𝑏62.9°𝑏𝑙32.9°\Delta\theta_{\mathrm{p}}^{\mathrm{g-e}}=\arctan(\frac{\cos(l-32.9\arcdeg)}{% \cos b\cot 62.9\arcdeg-\sin b\sin(l-32.9\arcdeg)})roman_Δ italic_θ start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_g - roman_e end_POSTSUPERSCRIPT = roman_arctan ( divide start_ARG roman_cos ( italic_l - 32.9 ° ) end_ARG start_ARG roman_cos italic_b roman_cot 62.9 ° - roman_sin italic_b roman_sin ( italic_l - 32.9 ° ) end_ARG ) (1)

is the angle between the galactic and equatorial reference directions (Corradi et al., 1998). For G28.34 at l=28.34°𝑙28.34°l=28.34\arcdegitalic_l = 28.34 ° and b=0.06°𝑏0.06°b=0.06\arcdegitalic_b = 0.06 °, we adopt Δθpge62.82°Δsuperscriptsubscript𝜃pge62.82°\Delta\theta_{\mathrm{p}}^{\mathrm{g-e}}\approx 62.82\arcdegroman_Δ italic_θ start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_g - roman_e end_POSTSUPERSCRIPT ≈ 62.82 °. The uncertainty on the Planck polarization position angle is given by δθ28°.65(δPI/PI)similar-to𝛿𝜃28°.65𝛿𝑃𝐼𝑃𝐼\delta\theta\sim 28\arcdeg.65(\delta PI/PI)italic_δ italic_θ ∼ 28 ° .65 ( italic_δ italic_P italic_I / italic_P italic_I ).

The Planck and JCMT observations are at the same frequency, so we could compare their intensities to investigate the consistency of the two datasets. To compare the Planck and JCMT data, we convolve the JCMT I𝐼Iitalic_I and PI𝑃𝐼PIitalic_P italic_I maps to the same resolution as the Planck maps. At 5\arcmin resolution, the JCMT peak intensity is 98 and 3.1 mJy beam11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT for I𝐼Iitalic_I and PI𝑃𝐼PIitalic_P italic_I, respectively. At the same position, the Planck intensity is 240 and 3.7 mJy beam11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT for I𝐼Iitalic_I and PI𝑃𝐼PIitalic_P italic_I, respectively. The JCMT peak I𝐼Iitalic_I and PI𝑃𝐼PIitalic_P italic_I values are similar-to\sim41% and similar-to\sim84% of the Planck values, respectively. This comparison suggests that the JCMT POL-2 data filters out a significant amount of the total intensity333POL-2 data reduction filters out the atmospheric emission as well as the extended emission for I𝐼Iitalic_I, Q𝑄Qitalic_Q, and U𝑈Uitalic_U maps. but recovers the majority of the polarized intensity.

2.4 FCRAO-14m 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO (1-0) data

The FCRAO-14m 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO (1-0) data are adopted from the Galactic Ring Survey (GRS; Jackson et al., 2006). The spatial and spectral resolutions of the 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO (1-0) data are similar-to\sim46\arcsec (similar-to\sim1 pc) and 0.21 km s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT, respectively. The pixel size is 22.14\arcsec. The typical sensitivity of the GRS survey is 0.13 K (in TAsubscriptsuperscript𝑇AT^{\ast}_{\mathrm{A}}italic_T start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_A end_POSTSUBSCRIPT). The main beam temperature is estimated from TAsubscriptsuperscript𝑇AT^{\ast}_{\mathrm{A}}italic_T start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_A end_POSTSUBSCRIPT adopting a main beam efficiency of 0.48. The 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO (1-0) data were previously reported in Beuther et al. (2020).

3 Results and analyses

3.1 Dust polarization and magnetic fields

Here we briefly overview the multi-scale magnetic field structures in G28.34 revealed by Planck, JCMT, and ALMA. Assuming that the observed linear polarization of the dust continuum is due to dust grain alignment, we rotate the dust polarization position angle by 90°°\arcdeg° to infer the magnetic field orientation.

Figure 1(a) shows the large-scale magnetic field orientation surrounding G28.34 revealed by Planck. The well-ordered large-scale magnetic field shows a predominant northeast-southwest orientation along the galactic plane.


Figure 1: (a). Magnetic field orientations revealed by Planck (black line segments) 0.85 mm dust polarization data overlaid on the Planck dust optical depth map (color scales) toward the IRDC G28.34 and its surrounding materials. Line segments are of arbitrary length. The 5\arcmin (similar-to\sim7 pc) beam (white circle) and a scale bar of 10 pc are indicated in the lower left and right corners, respectively. A white dashed line indicates the galactic plane (b=0°𝑏0°b=0\arcdegitalic_b = 0 °). The dashed rectangle indicates the JCMT map area in (b). (b). Magnetic field orientations revealed by JCMT POL-2 0.85 mm dust polarization observations overlaid on the JCMT 0.85 mm total intensity map (color scales. Only area with S/N(I𝐼Iitalic_I)>>>25 is shown) toward the IRDC G28.34. Black and grey line segments indicate S/N(PI𝑃𝐼PIitalic_P italic_I)>>>3 and 2<<<S/N(PI𝑃𝐼PIitalic_P italic_I)<<<3, respectively. Line segments are of arbitrary length. The 14\arcsec (similar-to\sim0.33 pc) beam (black circle) and a scale bar of 1 pc are indicated in the lower left and right corner, respectively. The infrared-bright and infrared-dark molecular clumps identified by Rathborne et al. (2006) are labeled with green and brown numbers, respectively. Red dashed contours indicate the ALMA fields of MM4 and MM9 corresponding to primary-beam responses of 0.5. Orange dashed circles mark the regions for MM4, MM6, MM9, and MM10 within which we use the polarization position angles to calculate the magnetic field strength in Section 3.4.2.

Figure 1(b) shows the magnetic field orientation in G28.34 and several nearby massive clumps revealed by JCMT. The magnetic field morphology in G28.34 is complex. Along the ridge of the main straight dark filament (containing MM10, MM14, MM4, and MM9), the magnetic field seems to be perpendicular to the filament spine, which is in agreement with previous observational studies of magnetic fields in filamentary IRDCs (Liu et al., 2018; Soam et al., 2019; Tang et al., 2019; Añez-López et al., 2020) and might be a result of gravitational accretion flows (Gómez et al., 2018; Li et al., 2018). In the northwestern clumps (MM1, MM2, MM11, and MM16) and in the diffuse region to the southeast of MM4 and MM9, the magnetic field tends to be parallel to the main straight dark filament, which may suggest that the magnetic field has kept its initial configuration inheriting from the large-scale magnetic field in these regions.

Figure 2 shows the magnetic field orientation in clumps MM4 and MM9 revealed by ALMA. Overall, our three configurations-combined (C-1, C-3, and C-4) ALMA images exhibit magnetic field morphologies similar to the two configurations-combined (C-1 and C-3) ALMA images reported by Liu et al. (2020).

Figure 2: Magnetic field orientations revealed by ALMA 1.3 mm dust polarization observations overlaid on the ALMA 1.3 mm total intensity map (color scales) toward MM4 and MM9. Black and grey line segments indicate S/N(PI𝑃𝐼PIitalic_P italic_I)>>>3 and 2<<<S/N(PI𝑃𝐼PIitalic_P italic_I)<<<3, respectively. Line segments are of arbitrary length. The 0.7×0.5similar-toabsent0.70.5\sim 0.7\arcsec\times 0.5\arcsec∼ 0.7 ″ × 0.5 ″ (similar-to\sim0.016-0.012 pc) synthesized beam (black ellipse) and a scale bar of 0.1 pc are indicated in the lower left and right corners, respectively.

3.1.1 Comparing multi-scale magnetic fields

One question to be addressed is how the small-scale magnetic field is correlated with the large-scale magnetic field (e.g., Li et al., 2009; Zhang et al., 2014; Li et al., 2015a). To study this, we first compare the JCMT magnetic field orientation (θJCMTsubscript𝜃JCMT\theta_{\mathrm{JCMT}}italic_θ start_POSTSUBSCRIPT roman_JCMT end_POSTSUBSCRIPT) in every pixel of the JCMT detection area (S/N(PI𝑃𝐼PIitalic_P italic_I)>>>2) with the Planck magnetic field orientation (θPlancksubscript𝜃Planck\theta_{\mathrm{Planck}}italic_θ start_POSTSUBSCRIPT roman_Planck end_POSTSUBSCRIPT) at the nearest pixel of the Planck map and calculate their angular difference with approaches similar to Zhang et al. (2014). Figure 3 overlays the magnetic field orientations revealed by Planck and JCMT. Figure 4(a) shows the spatial distribution of the absolute angular difference between the Planck and JCMT magnetic field orientations. Figure 5(a) shows the histogram of the absolute angular difference between the Planck and JCMT magnetic field orientations. Although overall there is no strong relation between the Planck and JCMT magnetic field orientations (Figure 5(a)), the spatial distribution for their angular difference is not random (as mentioned above and see Figure 4(a)). The difference between global and local statistics signifies the importance of investigating the local distributions of the angular difference. Thus, we further investigate the relation between the cloud- and clump-scale averaged magnetic fields (e.g., Li et al., 2009). To do this, we average the Planck polarization data (i.e., Q𝑄Qitalic_Q and U𝑈Uitalic_U) toward the cloud within a circle of 10 pc and the JCMT polarization detection toward each clump within a circle of 1 pc, and recalculate the polarization position angles. We find that half of the clumps (MM1, MM2, MM11, MM16, and MM17) within our studied region have averaged magnetic fields aligned within 30°°\arcdeg° of the cloud-scale magnetic field, while the other half of the clumps (MM4, MM6, MM9, MM10, and MM14) have averaged magnetic fields misaligned at 60°°\arcdeg°-90°°\arcdeg° with respect to the cloud-scale magnetic field (see Figure 5(b)). The bimodal distribution implies that the clump-scale magnetic field is organized with respect to the uniform cloud-scale Planck field. Otherwise, a randomly-orientated small-scale magnetic field cannot produce the observed bimodal distribution (Zhang et al., 2014). Future larger-sample observational studies are required to understand whether the bimodal distribution for angular differences between cloud- and clump-scale averaged magnetic fields is ubiquitous, and future theoretical and numerical studies will be needed to understand the physical mechanism behind this bimodal distribution.

Figure 3: Magnetic field orientations revealed by Planck (orange lines) and JCMT (white lines) overlaid on the Spitzer three-color composite image (red/green/blue = 8.0/4.5/3.6 μ𝜇\muitalic_μm). The Spitzer data is from the GLIMPSE project (Churchwell et al., 2009).

Similarly, we investigate the angular difference between the polarization position angles revealed by JCMT and ALMA. Figures 4(b) and 5(b) show the spatial distribution and histogram, respectively, for the angular difference between the JCMT and ALMA magnetic field orientations toward MM4. The local spatial distribution of the angular difference (Figure 4(b)) is complex and we refrain from describing them in detail. The histogram of angular difference (Figure 5(c)) shows that the magnetic field on scales of cores and condensations in MM4, as revealed by ALMA, is preferentially aligned with the clump-scale magnetic field revealed by JCMT. We further compare the averaged clump-scale and core-scale magnetic fields for core1--5 in MM4. To calculate the averaged magnetic fields of cores, we adopt the core positions in Zhang et al. (2009) and average the polarization detections of each core within a radius of 0.1 pc in a way similar to the derivation of the averaged clump-scale magnetic fields. We find that most cores (i.e., core1, 2, 4, and 5) have averaged magnetic fields aligned within 20°°\arcdeg° of the clump-scale magnetic field of MM4 (see Figure 5(d)). The preservation of magnetic field orientation could suggest that the magnetic field plays a crucial role in the collapse of this clump and in the formation of dense cores within (Li et al., 2015b). The clump-scale magnetic field is perpendicular to the chain of cores in MM4, which could suggest that the fragmentation in this clump is regulated by the magnetic field in the parental clump (Nakamura & Li, 2008). We refrain from comparing the averaged condensation-scale magnetic fields with large-scale magnetic fields in MM4 because the fragmentation in this clump at condensation level is complex (Zhang et al., 2015; Kong, 2019). On the other hand, the marginal polarization detection in MM9 does not support any statistics. With a rough comparison, we find that the angle between the clump-scale magnetic field of MM9 and the condensation-scale magnetic field in C1-Sa is similar-to\sim65°°\arcdeg°.

Figure 4: (a). Map of absolute angular differences (color scale) between magnetic field orientations revealed by Planck and JCMT in G28.34. The magnetic field orientation revealed by Planck is shown as grey line segments. The average magnetic field orientation within each 1-pc clump is indicated as black line segments. Contour starts at 50 mJy beam11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT and continues at 150 mJy beam11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT. (b). Map of absolute angular differences (color scale) between magnetic field orientations revealed by JCMT and ALMA in MM4. The magnetic field orientation revealed by JCMT is shown as grey line segments. The average magnetic field orientation within each 0.1-pc core is indicated as black line segments. Contour levels are (±plus-or-minus\pm±3, 6, 10, 20, 30, 40, 50, 70, 90, 110, 150, 180, 210, 250, 290, 340, 390, 450) ×σIabsentsubscript𝜎𝐼\times\sigma_{I}× italic_σ start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT.
Figure 5: (a). Histogram of angular differences between magnetic field orientations revealed by Planck and JCMT in G28.34. (b). Histogram of angular differences between averaged cloud- and clump-scale magnetic fields in G28.34. (c). Histogram of angular differences between magnetic field orientations revealed by JCMT and ALMA in MM4. (d). Histogram of angular differences between averaged clump- and core-scale magnetic fields in MM4.

3.2 Density and mass

Here we briefly describe the estimation of the gas density and mass from dust continuum and molecular line data. Then we investigate the mass-radius relation and density-radius relation.

3.2.1 Core

For the dense cores revealed by ALMA observations, we estimate the dust mass with

Mdust=Fνd2κνBν(T),subscript𝑀dustsubscript𝐹𝜈superscript𝑑2subscript𝜅𝜈subscript𝐵𝜈𝑇M_{\mathrm{dust}}=\frac{F_{\mathrm{\nu}}d^{2}}{\kappa_{\nu}B_{\nu}(T)},italic_M start_POSTSUBSCRIPT roman_dust end_POSTSUBSCRIPT = divide start_ARG italic_F start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_κ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT italic_B start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ( italic_T ) end_ARG , (2)

where Fνsubscript𝐹𝜈F_{\mathrm{\nu}}italic_F start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT is the flux density at frequency ν𝜈\nuitalic_ν, d𝑑ditalic_d is the distance, κν=(ν/1THz)βsubscript𝜅𝜈superscript𝜈1THz𝛽\kappa_{\nu}=(\nu/1\mathrm{THz})^{\beta}italic_κ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT = ( italic_ν / 1 roman_T roman_H roman_z ) start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT is the dust opacity (Hildebrand, 1983) in m22{}^{2}start_FLOATSUPERSCRIPT 2 end_FLOATSUPERSCRIPT kg11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT, and Bν(T)subscript𝐵𝜈𝑇B_{\nu}(T)italic_B start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ( italic_T ) is the Planck function at temperature T𝑇Titalic_T. We adopt a dust emissivity index (β𝛽\betaitalic_β) of similar-to\sim1.5 (e.g., Beuther et al., 2007; Chen et al., 2007) and a gas temperature of 15 K (Wang, 2018). Adopting a gas-to-dust ratio of Λ=100Λ100\Lambda=100roman_Λ = 100 (Savage & Jenkins, 1972), the gas mass can be estimated with Mgas=ΛMdustsubscript𝑀gasΛsubscript𝑀dustM_{\mathrm{gas}}=\Lambda M_{\mathrm{dust}}italic_M start_POSTSUBSCRIPT roman_gas end_POSTSUBSCRIPT = roman_Λ italic_M start_POSTSUBSCRIPT roman_dust end_POSTSUBSCRIPT. Then the gas column density is estimated with

NH2=MgasμH2mHA,subscript𝑁subscriptH2subscript𝑀gassubscript𝜇subscriptH2subscript𝑚H𝐴N_{\mathrm{H_{2}}}=\frac{M_{\mathrm{gas}}}{\mu_{\mathrm{H_{2}}}m_{\mathrm{H}}A},italic_N start_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT = divide start_ARG italic_M start_POSTSUBSCRIPT roman_gas end_POSTSUBSCRIPT end_ARG start_ARG italic_μ start_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_m start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT italic_A end_ARG , (3)

where μH2=2.8subscript𝜇subscriptH22.8\mu_{\mathrm{H_{2}}}=2.8italic_μ start_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT = 2.8 is the mean molecular weight per hydrogen molecule (Kauffmann et al., 2008), mHsubscript𝑚Hm_{\mathrm{H}}italic_m start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT is the atomic mass of hydrogen, and A𝐴Aitalic_A is the area.

3.2.2 Clump and cloud

We adopt the column density map (hereafter the COMB map) at intermediate scales from Lin et al. (2017). The COMB map was made by performing an iterative SED fitting procedure for multi-wavelength (70 μ𝜇\muitalic_μm to 850 μ𝜇\muitalic_μm) continuum data from Herschel, Caltech Submillimeter Observatory (CSO), JCMT, and Planck, using image combination techniques to recover extended emission for ground-based telescopes while preserving the high angular resolution. Details about basic data combinations and SED fitting procedures can be found in Lin et al. (2016). In particular, we used an updated combination procedure similar to Jiao et al. (2022), where the Planck map is first deconvolved based on an extrapolated model image. The updated combined image benefits from a better dynamical range to recover more extended emissions. The resolution of the resulting column density maps is 10\arcsec. The size of the COMB map is similar-to\sim600\arcsec (similar-to\sim7 pc). With the column density map, the gas mass is derived using Equation 3.

3.2.3 Environmental gas

For the environmental gas surrounding IRDC G28.34, we calculate the gas column density from the Planck τ353subscript𝜏353\tau_{353}italic_τ start_POSTSUBSCRIPT 353 end_POSTSUBSCRIPT map and the FCRAO-14m 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO (1-0) data.

For the Planck data, we convert τ353subscript𝜏353\tau_{353}italic_τ start_POSTSUBSCRIPT 353 end_POSTSUBSCRIPT to the column density of hydrogen atoms (NHsubscript𝑁HN_{\mathrm{H}}italic_N start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT) with the relation (Planck Collaboration et al., 2014, 2016)

τ353/NH=1.2×1026cm2.subscript𝜏353subscript𝑁H1.2superscript1026superscriptcm2\tau_{353}/N_{\mathrm{H}}=1.2\times 10^{-26}\mathrm{cm}^{2}.italic_τ start_POSTSUBSCRIPT 353 end_POSTSUBSCRIPT / italic_N start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT = 1.2 × 10 start_POSTSUPERSCRIPT - 26 end_POSTSUPERSCRIPT roman_cm start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT . (4)

Note that NHsubscript𝑁HN_{\mathrm{H}}italic_N start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT accounts for the column density of both the atomic gas and molecular gas (Planck Collaboration et al., 2014, 2016) along the line of sight (LOS). The gas mass is given by

Mgas=μHmHANH,subscript𝑀gassubscript𝜇Hsubscript𝑚H𝐴subscript𝑁HM_{\mathrm{gas}}=\mu_{\mathrm{H}}m_{\mathrm{H}}AN_{\mathrm{H}},italic_M start_POSTSUBSCRIPT roman_gas end_POSTSUBSCRIPT = italic_μ start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT italic_m start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT italic_A italic_N start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT , (5)

where we assume the mean atomic weight per hydrogen atom is μH1.4similar-tosubscript𝜇H1.4\mu_{\mathrm{H}}\sim 1.4italic_μ start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT ∼ 1.4.

For the 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO (1-0) data, we estimate the integrated line intensity with W13CO=ΣiNchIiΔvchsubscript𝑊13COsuperscriptsubscriptΣ𝑖subscript𝑁chsubscript𝐼𝑖Δsubscript𝑣chW_{\mathrm{13CO}}=\Sigma_{i}^{N_{\mathrm{ch}}}I_{i}\Delta v_{\mathrm{ch}}italic_W start_POSTSUBSCRIPT 13 roman_C roman_O end_POSTSUBSCRIPT = roman_Σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT roman_ch end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_I start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT roman_Δ italic_v start_POSTSUBSCRIPT roman_ch end_POSTSUBSCRIPT, where Iisubscript𝐼𝑖I_{i}italic_I start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, ΔvchΔsubscript𝑣ch\Delta v_{\mathrm{ch}}roman_Δ italic_v start_POSTSUBSCRIPT roman_ch end_POSTSUBSCRIPT, and Nchsubscript𝑁chN_{\mathrm{ch}}italic_N start_POSTSUBSCRIPT roman_ch end_POSTSUBSCRIPT are the line intensity, channel width, and the number of integrated channels, respectively. The 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO (1-0) line is integrated between 71 and 86 km s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT to cover the main velocity component of the cloud (Beuther et al., 2020). Because the 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO (1-0) line peak brightness temperature is much smaller than the gas temperature, this line is very likely optically thin (Beuther et al., 2020). In the optically thin case, the upper state column density is given by (Goldsmith & Langer, 1999)

Nu=8πkBν2W13COhc3Aul,subscript𝑁𝑢8𝜋subscript𝑘𝐵superscript𝜈2subscript𝑊13COsuperscript𝑐3subscript𝐴𝑢𝑙N_{u}=\frac{8\pi k_{B}\nu^{2}W_{\mathrm{13CO}}}{hc^{3}A_{ul}},italic_N start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT = divide start_ARG 8 italic_π italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_ν start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_W start_POSTSUBSCRIPT 13 roman_C roman_O end_POSTSUBSCRIPT end_ARG start_ARG italic_h italic_c start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_A start_POSTSUBSCRIPT italic_u italic_l end_POSTSUBSCRIPT end_ARG , (6)

where kBsubscript𝑘𝐵k_{B}italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT, hhitalic_h, and Aulsubscript𝐴𝑢𝑙A_{ul}italic_A start_POSTSUBSCRIPT italic_u italic_l end_POSTSUBSCRIPT are the Boltzmann constant, Planck constant, and Einstein coefficient, respectively. Assuming local thermal equilibrium, the total column density of 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO is given by

NC13O=NuZgueEu/kBT,subscript𝑁superscript𝐶13𝑂subscript𝑁u𝑍subscript𝑔usuperscript𝑒subscript𝐸usubscript𝑘𝐵𝑇N_{{}^{13}CO}=\frac{N_{\mathrm{u}}Z}{g_{\mathrm{u}}e^{-E_{\mathrm{u}}/k_{B}T}},italic_N start_POSTSUBSCRIPT start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPT italic_C italic_O end_POSTSUBSCRIPT = divide start_ARG italic_N start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT italic_Z end_ARG start_ARG italic_g start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_E start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT / italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T end_POSTSUPERSCRIPT end_ARG , (7)

where gusubscript𝑔ug_{\mathrm{u}}italic_g start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT, Eusubscript𝐸uE_{\mathrm{u}}italic_E start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT, and Z𝑍Zitalic_Z are the statistical weight of the upper state, the upper energy level, and the partition function, respectively. Here we adopt T=15𝑇15T=15italic_T = 15 K as well. The values of Aulsubscript𝐴𝑢𝑙A_{ul}italic_A start_POSTSUBSCRIPT italic_u italic_l end_POSTSUBSCRIPT, gusubscript𝑔ug_{\mathrm{u}}italic_g start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT, Eusubscript𝐸uE_{\mathrm{u}}italic_E start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT, and Z𝑍Zitalic_Z are adopted from the CDMS (Müller et al., 2001) and LAMDA (Schöier et al., 2005) databases. Thus, the column density of H22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT can be estimated using the standard abundance (Wilson et al., 2013)

NH2=NC13O×4.6×105.subscript𝑁subscript𝐻2subscript𝑁superscript𝐶13𝑂4.6superscript105N_{H_{2}}=N_{{}^{13}CO}\times 4.6\times 10^{5}.italic_N start_POSTSUBSCRIPT italic_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT = italic_N start_POSTSUBSCRIPT start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPT italic_C italic_O end_POSTSUBSCRIPT × 4.6 × 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT . (8)

The standard abundance is only valid when NH2<5×1021subscript𝑁subscript𝐻25superscript1021N_{H_{2}}<5\times 10^{21}italic_N start_POSTSUBSCRIPT italic_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT < 5 × 10 start_POSTSUPERSCRIPT 21 end_POSTSUPERSCRIPT cm22{}^{-2}start_FLOATSUPERSCRIPT - 2 end_FLOATSUPERSCRIPT and could present a scatter of factor two to five (Wilson et al., 2013). Because CO isotopes could be depleted in dark clouds (Bergin & Tafalla, 2007), NH2subscript𝑁subscript𝐻2N_{H_{2}}italic_N start_POSTSUBSCRIPT italic_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT toward IRDC G28.34 estimated with 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO may only be a lower limit. With the estimated gas column density, the gas mass is derived with Equation 3.

We find that the ratio between 0.5NH0.5subscript𝑁H0.5N_{\mathrm{H}}0.5 italic_N start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT estimated with the Planck τ353subscript𝜏353\tau_{353}italic_τ start_POSTSUBSCRIPT 353 end_POSTSUBSCRIPT map and NH2subscript𝑁subscript𝐻2N_{H_{2}}italic_N start_POSTSUBSCRIPT italic_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT estimated with the 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO data is similar-to\sim10 toward the center of G28.34. Their ratio gradually increases to similar-to\sim1033{}^{3}start_FLOATSUPERSCRIPT 3 end_FLOATSUPERSCRIPT toward the edge of the Planck map. As the Planck observations trace all the atomic and molecular gases along the LOS, it is reasonable that the column density traced by Planck should be higher than the molecular gas integrated within a specific velocity range. On the other hand, the systematic lower gas column densities of 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO might be due to the 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO depletion, the FCRAO-14m observation filtering out the large-scale emission at the extent of the off-position, and the uncertainties in the adopted 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO abundance.

3.2.4 Scaling relations

The mass-radius (Mr𝑀𝑟M-ritalic_M - italic_r) and density-radius (nr𝑛𝑟n-ritalic_n - italic_r) relations are important properties of star formation regions (Larson, 1981). With the estimated column density and gas mass, we investigate the Mr𝑀𝑟M-ritalic_M - italic_r and nr𝑛𝑟n-ritalic_n - italic_r relations for the cores and clumps in IRDC G28.34 and in its surrounding gas.

With the gas mass derived from the Planck τ353subscript𝜏353\tau_{353}italic_τ start_POSTSUBSCRIPT 353 end_POSTSUBSCRIPT map and from the FCRAO-14m 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO (1-0) data, we obtain the mass profile for the environmental gas within circles of different radii (from beam size to 20 pc) centered at G28.34. With the COMB mass map, we obtain the mass profile for each infrared dark clump within circles of different radii (from beam size to 2 pc) centered at the local continuum peak. For simplicity, we only consider clumps with both molecular line detection and sufficient dust polarization detection (i.e., MM4, MM6, MM9, MM10, and MM14) in our analysis. MM10 and MM14 are unseparated at the resolution of the COMB map and are thus considered as one clump in our analysis. We also obtain a mass estimation for the whole COMB map area (rsimilar-to𝑟absentr\simitalic_r ∼7 pc). At the resolution of ALMA, the fragmentation status is complicated in MM4 and MM9. Thus, we only report the total mass and effective radius (A/π𝐴𝜋\sqrt{A/\pi}square-root start_ARG italic_A / italic_π end_ARG) for the ALMA area with S/N(I𝐼Iitalic_I)>>>5. Similarly, we obtain the profile for the average column density within circles of different radii. Assuming the studied structures are spherical, the average number density is estimated from the average column density with n=0.75N/r𝑛0.75𝑁𝑟n=0.75N/ritalic_n = 0.75 italic_N / italic_r.

Figure 6: (a). Mass-radius relation. The red solid line indicates the empirical threshold for massive star formation (Mlim=870M(r/pc)1.33subscript𝑀𝑙𝑖𝑚870subscript𝑀direct-productsuperscript𝑟pc1.33M_{lim}=870M_{\odot}(r/\mathrm{pc})^{1.33}italic_M start_POSTSUBSCRIPT italic_l italic_i italic_m end_POSTSUBSCRIPT = 870 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT ( italic_r / roman_pc ) start_POSTSUPERSCRIPT 1.33 end_POSTSUPERSCRIPT, Kauffmann et al., 2010). The blue solid line indicates the Larson’s third law (MLarson=460M(r/pc)1.9subscript𝑀𝐿𝑎𝑟𝑠𝑜𝑛460subscript𝑀direct-productsuperscript𝑟𝑝𝑐1.9M_{Larson}=460M_{\odot}(r/pc)^{1.9}italic_M start_POSTSUBSCRIPT italic_L italic_a italic_r italic_s italic_o italic_n end_POSTSUBSCRIPT = 460 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT ( italic_r / italic_p italic_c ) start_POSTSUPERSCRIPT 1.9 end_POSTSUPERSCRIPT, Larson, 1981). (b). Average number density-radius relation. The blue solid line indicates the Larson’s third law (nLarson=1586subscript𝑛𝐿𝑎𝑟𝑠𝑜𝑛1586n_{Larson}=1586italic_n start_POSTSUBSCRIPT italic_L italic_a italic_r italic_s italic_o italic_n end_POSTSUBSCRIPT = 1586 cm(r/pc)1.13{}^{-3}(r/\mathrm{pc})^{-1.1}start_FLOATSUPERSCRIPT - 3 end_FLOATSUPERSCRIPT ( italic_r / roman_pc ) start_POSTSUPERSCRIPT - 1.1 end_POSTSUPERSCRIPT, Larson, 1981). Different colors represent different regions. Different symbols indicate different datasets. The black dashed lines present the results of a simple power-law fit for the ALMA data and the COMB data.

Figure 6(a) shows the obtained mass-radius relation. An obvious trend is that the mass and radius for the ALMA data and the COMB data are positively correlated. With a simple minimum chi-squared power-law fit, we obtain Mr1.59±0.14proportional-to𝑀superscript𝑟plus-or-minus1.590.14M\propto r^{1.59\pm 0.14}italic_M ∝ italic_r start_POSTSUPERSCRIPT 1.59 ± 0.14 end_POSTSUPERSCRIPT. Here we assume an uncertainty of a factor of 2 for the mass during the fitting. The ALMA data for MM4 is higher than the fitted power law, which might be due to the deviation from the assumed spherical structure in this clump. The fitted power law is higher than the empirical threshold for massive star formation (Mlim=870M(r/pc)1.33subscript𝑀𝑙𝑖𝑚870subscript𝑀direct-productsuperscript𝑟𝑝𝑐1.33M_{lim}=870M_{\odot}(r/pc)^{1.33}italic_M start_POSTSUBSCRIPT italic_l italic_i italic_m end_POSTSUBSCRIPT = 870 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT ( italic_r / italic_p italic_c ) start_POSTSUPERSCRIPT 1.33 end_POSTSUPERSCRIPT, Kauffmann et al., 2010), indicating that IRDC G28.34 is potentially forming massive stars. The obtained power-law index of 1.59 is smaller than the value of 1.9 reported by Larson (1981). Note that our Mr𝑀𝑟M-ritalic_M - italic_r relation presents the relation in an individual cloud, while Larson’s third law presents the relation for an ensemble of cloud samples, so the two relations may not be comparable. The mass derived from the Planck map is higher than the fitted power law, while the mass derived from the FCRAO-14m 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO data is lower than the fitted power law. As discussed in Section 3.2.3, this deviation might be because the molecular mass is overestimated by the Planck map and underestimated by the FCRAO-14m 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO data.

Figure 6(b) shows the obtained average number density-radius relation. With a power-law fit for the ALMA data and the COMB data, we obtain nr1.41±0.14proportional-to𝑛superscript𝑟plus-or-minus1.410.14n\propto r^{-1.41\pm 0.14}italic_n ∝ italic_r start_POSTSUPERSCRIPT - 1.41 ± 0.14 end_POSTSUPERSCRIPT. Here we assume an uncertainty of a factor of 2 for the density during the fitting. The obtained power-law index of -1.41 is steeper than the value of -1.1 reported by Larson (1981). Similarly, our nr𝑛𝑟n-ritalic_n - italic_r relation in an individual cloud may not be comparable to Larson’s third law which presents the relation for an ensemble of cloud samples.

3.3 Molecular line and velocity fields

Here we briefly overview the multi-scale velocity structures revealed by the FCRAO-14m 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO (1-0), JCMT 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO (3-2) and HCO+{}^{+}start_FLOATSUPERSCRIPT + end_FLOATSUPERSCRIPT (4-3), and ALMA N22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPTD+{}^{+}start_FLOATSUPERSCRIPT + end_FLOATSUPERSCRIPT (3-2) line data. The information of each line is summarised in Table 1. The value of line parameters are adopted from the CDMS (Müller et al., 2001) and LAMDA (Schöier et al., 2005) databases. No information on the collision rate coefficient exists for N22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPTD+{}^{+}start_FLOATSUPERSCRIPT + end_FLOATSUPERSCRIPT in LAMDA, so we adopt the values for N22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPTH+{}^{+}start_FLOATSUPERSCRIPT + end_FLOATSUPERSCRIPT instead. A particular line tracer is only sensitive to gas with densities above its critical density (ncsubscript𝑛cn_{\mathrm{c}}italic_n start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT) but no more than a factor of 2 orders of magnitude (Goodman et al., 1998). We calculate the intensity-weighted velocity centroid Vc(𝒙)subscript𝑉𝑐𝒙V_{c}(\boldsymbol{x})italic_V start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( bold_italic_x ) at position 𝒙𝒙\boldsymbol{x}bold_italic_x with

Vc(𝒙)=ΣiNchIi(𝒙)viΔvchΣiNchIi(𝒙)Δvch,subscript𝑉𝑐𝒙superscriptsubscriptΣ𝑖subscript𝑁chsubscript𝐼𝑖𝒙subscript𝑣𝑖Δsubscript𝑣chsuperscriptsubscriptΣ𝑖subscript𝑁chsubscript𝐼𝑖𝒙Δsubscript𝑣chV_{c}(\boldsymbol{x})=\frac{\Sigma_{i}^{N_{\mathrm{ch}}}I_{i}(\boldsymbol{x})v% _{i}\Delta v_{\mathrm{ch}}}{\Sigma_{i}^{N_{\mathrm{ch}}}I_{i}(\boldsymbol{x})% \Delta v_{\mathrm{ch}}},italic_V start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( bold_italic_x ) = divide start_ARG roman_Σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT roman_ch end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_I start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( bold_italic_x ) italic_v start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT roman_Δ italic_v start_POSTSUBSCRIPT roman_ch end_POSTSUBSCRIPT end_ARG start_ARG roman_Σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT roman_ch end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_I start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( bold_italic_x ) roman_Δ italic_v start_POSTSUBSCRIPT roman_ch end_POSTSUBSCRIPT end_ARG , (9)

where Ii(𝒙)subscript𝐼𝑖𝒙I_{i}(\boldsymbol{x})italic_I start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( bold_italic_x ), visubscript𝑣𝑖v_{i}italic_v start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, ΔvchΔsubscript𝑣ch\Delta v_{\mathrm{ch}}roman_Δ italic_v start_POSTSUBSCRIPT roman_ch end_POSTSUBSCRIPT, and Nchsubscript𝑁chN_{\mathrm{ch}}italic_N start_POSTSUBSCRIPT roman_ch end_POSTSUBSCRIPT are the line intensity, line-of-sight velocity, velocity channel width, and number of channels, respectively. The propagated uncertainty of Vc(𝒙)subscript𝑉𝑐𝒙V_{c}(\boldsymbol{x})italic_V start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( bold_italic_x ) is

δVc(𝒙)=σchΔvchΣiNch(viVc(𝒙))2ΣiNchIi(𝒙)Δvch,𝛿subscript𝑉𝑐𝒙subscript𝜎𝑐Δsubscript𝑣chsuperscriptsubscriptΣ𝑖subscript𝑁chsuperscriptsubscript𝑣𝑖subscript𝑉𝑐𝒙2superscriptsubscriptΣ𝑖subscript𝑁chsubscript𝐼𝑖𝒙Δsubscript𝑣ch\delta V_{c}(\boldsymbol{x})=\frac{\sigma_{ch}\Delta v_{\mathrm{ch}}\sqrt{% \Sigma_{i}^{N_{\mathrm{ch}}}(v_{i}-V_{c}(\boldsymbol{x}))^{2}}}{\Sigma_{i}^{N_% {\mathrm{ch}}}I_{i}(\boldsymbol{x})\Delta v_{\mathrm{ch}}},italic_δ italic_V start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( bold_italic_x ) = divide start_ARG italic_σ start_POSTSUBSCRIPT italic_c italic_h end_POSTSUBSCRIPT roman_Δ italic_v start_POSTSUBSCRIPT roman_ch end_POSTSUBSCRIPT square-root start_ARG roman_Σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT roman_ch end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ( italic_v start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_V start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( bold_italic_x ) ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG start_ARG roman_Σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT roman_ch end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_I start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( bold_italic_x ) roman_Δ italic_v start_POSTSUBSCRIPT roman_ch end_POSTSUBSCRIPT end_ARG , (10)

where σchsubscript𝜎𝑐\sigma_{ch}italic_σ start_POSTSUBSCRIPT italic_c italic_h end_POSTSUBSCRIPT is the noise of one spectral channel (see Section 2). For FCRAO-14m 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO (1-0), we only consider line emissions from 71 to 86 km s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT which covers the main velocity component of G28.34 at large scales (Beuther et al., 2020). There are no apparent LOS distant gas structures at this velocity superposed at the same POS position for this IRDC (Simon et al., 2006; Beuther et al., 2020). For JCMT observations toward the cloud, we only consider velocities within similar-to\sim5 km s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT with respect to the local-standard-of-rest (LSR) velocity (Vlsr79similar-tosubscript𝑉lsr79V_{\mathrm{lsr}}\sim 79italic_V start_POSTSUBSCRIPT roman_lsr end_POSTSUBSCRIPT ∼ 79 km s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT) of G28.34 to avoid potential contamination from high-velocity outflows. For ALMA observations toward MM4 and MM9, we only consider velocities within similar-to\sim3 km s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT with respect to the Vlsrsubscript𝑉lsrV_{\mathrm{lsr}}italic_V start_POSTSUBSCRIPT roman_lsr end_POSTSUBSCRIPT of each clump (similar-to\sim79 and 80 km s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT for MM4 and MM9, respectively, Zhang et al., 2015) because there is nearly no line emission at >>>3 km s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT. All lines are likely to be optically thin as the main-beam temperature of the line peaks is smaller than the gas/dust temperature of similar-to\sim15 K (see Appendix B).

Table 1: Summary of molecular line data
Line Frequency Eu/ksubscript𝐸u𝑘E_{\mathrm{u}}/kitalic_E start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT / italic_k aaUpper energy level in units of K. ncsubscript𝑛cn_{\mathrm{c}}italic_n start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT bbCritical density at 15 K. For N22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPTD+{}^{+}start_FLOATSUPERSCRIPT + end_FLOATSUPERSCRIPT, we adopt the collision rate coefficient of N22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPTH+{}^{+}start_FLOATSUPERSCRIPT + end_FLOATSUPERSCRIPT. Instrument lbeamsubscript𝑙beaml_{\mathrm{beam}}italic_l start_POSTSUBSCRIPT roman_beam end_POSTSUBSCRIPT ccResolution. lMRSsubscript𝑙MRSl_{\mathrm{MRS}}italic_l start_POSTSUBSCRIPT roman_MRS end_POSTSUBSCRIPT ddMaximum recoverable scale for ALMA. Targets
(GHz) (K) (cm33{}^{-3}start_FLOATSUPERSCRIPT - 3 end_FLOATSUPERSCRIPT) (\arcsec) (\arcsec)
N22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPTD+{}^{+}start_FLOATSUPERSCRIPT + end_FLOATSUPERSCRIPT (3-2) 231.3218 22.2 1.0 ×\times× 1066{}^{6}start_FLOATSUPERSCRIPT 6 end_FLOATSUPERSCRIPT ALMA 0.7 13 MM4,MM9
HCO+{}^{+}start_FLOATSUPERSCRIPT + end_FLOATSUPERSCRIPT (4-3) 356.7342 42.8 8.4 ×\times× 1066{}^{6}start_FLOATSUPERSCRIPT 6 end_FLOATSUPERSCRIPT JCMT 14 G28.34
1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO (3-2) 330.5880 31.7 2.9 ×\times× 1044{}^{4}start_FLOATSUPERSCRIPT 4 end_FLOATSUPERSCRIPT JCMT 14 G28.34
1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO (1-0) 110.2014 5.3 1.9 ×\times× 1033{}^{3}start_FLOATSUPERSCRIPT 3 end_FLOATSUPERSCRIPT FCRAO-14m 46
footnotetext: Notes. The line information is from the CDMS (Müller et al., 2001) and LAMDA (Schöier et al., 2005) databases.

3.3.1 Velocity centroid maps

Figures 7 and 8 show the multi-scale velocity structure surrounding and within G28.34. The FCRAO-14m 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO (1-0) data (see Figure 7(a)) reveals a velocity gradient across Galactic latitudes near G28.34 (Beuther et al., 2020). The origin of this global gradient is still unclear. At higher resolution, the JCMT 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO (3-2) data reveals a velocity gradient perpendicular to the main dark filament (see Figure 7(b)) that is consistent with the global velocity gradient. Previous studies have interpreted this gradient as gas flows converging at the position of the filament (Beuther et al., 2020). The origin of the converging flow may be related to gravitationally driven collapse or external compression (Beuther et al., 2020). The HCO+{}^{+}start_FLOATSUPERSCRIPT + end_FLOATSUPERSCRIPT (4-3) line with a higher critical density than the 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO (3-2) transition reveals the velocity structure of most clumps in G28.34 except for MM17 (see Figure 7(c)). Within the dark filament, the higher-density gas traced by the HCO+{}^{+}start_FLOATSUPERSCRIPT + end_FLOATSUPERSCRIPT (4-3) line is more redshifted than the lower-density gas traced by the 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO (3-2) transition. With the NH33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT observations, Zhang et al. (2015) have identified a longitudinal velocity gradient of 0.6 km s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT along the main dark filament from MM10 to MM14 to MM4, which reverses from MM4 to MM9. This velocity gradient is also seen in the JCMT 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO (3-2) and HCO+{}^{+}start_FLOATSUPERSCRIPT + end_FLOATSUPERSCRIPT (4-3) data. Toward MM1, there seems to be a north-south velocity gradient deviating from the trend of the large-scale gradient, which may be an indicator of local gravitational infall in this clump. At smaller scales, the variation of Vcsubscript𝑉𝑐V_{c}italic_V start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT is overall small in dense cores in MM4 and MM9 as seen by the ALMA N22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPTD+{}^{+}start_FLOATSUPERSCRIPT + end_FLOATSUPERSCRIPT (3-2) observations. There is a redshifted velocity component toward the streamer-like continuum structure in the north of MM4-core5, which may indicate that the core is accreting gas from the clump gas through the streamer. There is a possible sign of a redshifted velocity component in the north of C1-N. Other than these redshifted velocity components, there are no apparent signs of ordered velocity gradients in the major part of cores in MM4 and MM9.

Figure 7: (a) Velocity centroid map of the FCRAO-14m 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO (1-0) data in the surrounding of G28.34. Black contours correspond to the Planck τ353subscript𝜏353\tau_{353}italic_τ start_POSTSUBSCRIPT 353 end_POSTSUBSCRIPT map, starting from 0.0005 and continuing with an interval of 0.0005. A dashed line indicates the galactic plane (b=0°𝑏0°b=0\arcdegitalic_b = 0 °). The dashed rectangle indicates the JCMT map area in (b) and (c). (b)-(c) Velocity centroid maps of the 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO (3-2) and HCO+{}^{+}start_FLOATSUPERSCRIPT + end_FLOATSUPERSCRIPT (4-3) emission toward the IRDC G28.34 obtained with JCMT. Black contours correspond to the JCMT 0.85 mm dust continuum map. Contours start at 50 mJy beam11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT and continue at a step of 150 mJy beam11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT. The beam and a scale bar are indicated in the lower left and lower right corner of each panel, respectively.

Figure 8: Velocity centroid map of the ALMA N22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPTD+{}^{+}start_FLOATSUPERSCRIPT + end_FLOATSUPERSCRIPT (3-2) data toward MM4 and MM9. Contours correspond to the ALMA 1.3 mm dust continuum map. Contour levels are (±plus-or-minus\pm±3, 6, 10, 20, 30, 40, 50, 70, 90, 110, 150, 180, 210, 250, 290, 340, 390, 450) ×σIabsentsubscript𝜎𝐼\times\sigma_{I}× italic_σ start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT. The synthesized beam and a scale bar are indicated in the lower left and lower right corner of each panel, respectively.

3.3.2 Multi-scale statistics

Here we use the velocity dispersion-radius (σvrsubscript𝜎𝑣𝑟\sigma_{v}-ritalic_σ start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT - italic_r) relation and velocity centroid dispersion-radius (σcrsubscript𝜎𝑐𝑟\sigma_{c}-ritalic_σ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT - italic_r) relation to explore the multi-scale velocity statistics in G28.34.

In a similar way to the derivation of the Mr𝑀𝑟M-ritalic_M - italic_r relation (Section 3.2.4), we derive the σvrsubscript𝜎𝑣𝑟\sigma_{v}-ritalic_σ start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT - italic_r relation for the average spectra (see Appendix B) of each line within circles of different radii, except for the ALMA data where the spectra are averaged within the whole emission area. Similar to the approaches that most previous studies have adopted, we fit a single-component Gaussian profile for each line to obtain the Full Width at Half Maximum (FWHM. i.e., linewidth). The relation between the FWHM and velocity dispersion is FWHM=8ln2σv𝐹𝑊𝐻𝑀82subscript𝜎𝑣FWHM=\sqrt{8\ln{2}}\sigma_{v}italic_F italic_W italic_H italic_M = square-root start_ARG 8 roman_ln 2 end_ARG italic_σ start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT. For the JCMT data, we focus on the infrared dark clumps MM4, MM6, MM9, and MM10 (+MM14). The thermal velocity linewidth (<<<0.06 km s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT for our studied molecules) is much smaller than the non-thermal linewidth in our studied regions due to the low temperature, so we regard the total linewidth as the non-thermal linewidth and do not subtract the thermal contribution. It should be noted that the radius of each considered region does not necessarily reflect the corresponding spatial scale of the measured linewidth. This is because the linewidth is also dependent on the LOS depth, which cannot be easily determined from observations. As we only fit a single-component Gaussian profile for all the lines, the linewidth in some regions could be broadened by the superposition of substructures with different LOS local-standard-of-rest (LSR) velocities at small scales (i.e., indistinguishable multiple velocity components in the line profiles. See Appendix B). Due to these effects, the measured linewidth from the observed spectra likely presents the upper limit of the true non-thermal linewidth at the corresponding scale of the radius.

Figure 9: (a). Velocity dispersion-radius relation. The blue solid line indicates Larson’s first law (σ1D,Larson=0.83subscript𝜎1𝐷𝐿𝑎𝑟𝑠𝑜𝑛0.83\sigma_{1D,Larson}=0.83italic_σ start_POSTSUBSCRIPT 1 italic_D , italic_L italic_a italic_r italic_s italic_o italic_n end_POSTSUBSCRIPT = 0.83 km s(r/pc)0.381{}^{-1}(r/pc)^{0.38}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT ( italic_r / italic_p italic_c ) start_POSTSUPERSCRIPT 0.38 end_POSTSUPERSCRIPT, Larson, 1981). The original 3D velocity dispersion in Larson’s first law is converted to 1D velocity dispersion with the relation σ3D,Larson=3σ1D,Larsonsubscript𝜎3𝐷𝐿𝑎𝑟𝑠𝑜𝑛3subscript𝜎1𝐷𝐿𝑎𝑟𝑠𝑜𝑛\sigma_{3D,Larson}=\sqrt{3}\sigma_{1D,Larson}italic_σ start_POSTSUBSCRIPT 3 italic_D , italic_L italic_a italic_r italic_s italic_o italic_n end_POSTSUBSCRIPT = square-root start_ARG 3 end_ARG italic_σ start_POSTSUBSCRIPT 1 italic_D , italic_L italic_a italic_r italic_s italic_o italic_n end_POSTSUBSCRIPT. (b). Velocity centroid dispersion-radius relation. Only data points with l𝑙litalic_l greater than the beam resolution and smaller than the effective radius of the considered region are shown. Different colors represent different regions. Different symbols indicate different instruments and lines.

Figure 9(a) shows the velocity dispersion-radius relation. At first glimpse, it is obvious that the σvrsubscript𝜎𝑣𝑟\sigma_{v}-ritalic_σ start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT - italic_r relations from different instruments and lines are not continuous. The discontinuity of σvrsubscript𝜎𝑣𝑟\sigma_{v}-ritalic_σ start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT - italic_r relation is similar to what Peretto et al. (2023) have recently found in a large sample of 27 infrared dark clouds. While Peretto et al. (2023) interpreted the discontinuous σvrsubscript𝜎𝑣𝑟\sigma_{v}-ritalic_σ start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT - italic_r relation as clump dynamics decoupling from those of the parental clouds, we suggest that the discontinuity is more likely due to the fact that the physical conditions traced by different line data are different and that the σvrsubscript𝜎𝑣𝑟\sigma_{v}-ritalic_σ start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT - italic_r relation does not correctly present the true velocity dispersion-size relation of the gas in 3D.

As seen in Figure 9(a), the σvrsubscript𝜎𝑣𝑟\sigma_{v}-ritalic_σ start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT - italic_r relation for each region seems to be relatively flat with small slopes or even negative slopes for the single-dish (FCRAO-14m and JCMT) observations. The flat slopes suggest that our measured velocity dispersion is likely dominated by the LOS integration up to the largest depth of the gas traced by specific lines, which is similar to what was found in the low-mass cloud Polaris Flare (Ossenkopf & Mac Low, 2002). The JCMT velocity dispersions are lower than the FCRAO-14m 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO (1-0) velocity dispersions. This might be due to the higher ncsubscript𝑛𝑐n_{c}italic_n start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT of the 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO (3-2) and HCO+{}^{+}start_FLOATSUPERSCRIPT + end_FLOATSUPERSCRIPT (4-3) lines, which trace fewer substructures along the LOS. On the other hand, the single-dish observations filter out the diffuse emission at the extent of the off position. The particular JCMT observations may have set a nearer off position than the FCRAO-14m observations, which could filter out more diffuse emissions and is also plausible to account for the lower velocity dispersions revealed by JCMT. For the JCMT observations, the velocity dispersions of HCO+{}^{+}start_FLOATSUPERSCRIPT + end_FLOATSUPERSCRIPT (4-3) are lower than that traced by 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO (3-2), which might be due to the higher ncsubscript𝑛𝑐n_{c}italic_n start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT and fewer LOS substructures traced by HCO+{}^{+}start_FLOATSUPERSCRIPT + end_FLOATSUPERSCRIPT (4-3). In addition, the ALMA velocity dispersions are much lower than those traced by JCMT and FCRAO-14m, which may mainly be because the emissions from lower-density substructures are filtered out by the interferometric observations.

We also study the scaling relation of velocity fields with the dispersion (i.e., standard deviation) of velocity centroids (σcsubscript𝜎𝑐\sigma_{c}italic_σ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT). The velocity centroid is insensitive to thermal broadening and thus primarily reveals non-thermal motions (Lazarian et al., 2022). Moreover, velocity centroids present averaged statistics along the LOS and are not affected by the broadening effect of LOS depth as for the linewidth. Thus the radius in the σcrsubscript𝜎𝑐𝑟\sigma_{c}-ritalic_σ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT - italic_r relation correctly reflects the true scale in the scaling relation. Due to the LOS averaging, for a single-component turbulent structure, the turbulent velocity centroid dispersion is smaller than the turbulent velocity dispersion by a factor of σv/σc=Nturb=L/λcsubscript𝜎𝑣subscript𝜎𝑐subscript𝑁𝑡𝑢𝑟𝑏𝐿subscript𝜆𝑐\sigma_{v}/\sigma_{c}=\sqrt{N_{turb}}=\sqrt{L/\lambda_{c}}italic_σ start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT / italic_σ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = square-root start_ARG italic_N start_POSTSUBSCRIPT italic_t italic_u italic_r italic_b end_POSTSUBSCRIPT end_ARG = square-root start_ARG italic_L / italic_λ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG (Dickman & Kleiner, 1985), where Nturbsubscript𝑁𝑡𝑢𝑟𝑏N_{turb}italic_N start_POSTSUBSCRIPT italic_t italic_u italic_r italic_b end_POSTSUBSCRIPT is the number of independent turbulent cells along the LOS, L𝐿Litalic_L is the effective depth, and λcsubscript𝜆𝑐\lambda_{c}italic_λ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT is the turbulent correlation length. Previous observational Vcsubscript𝑉𝑐V_{c}italic_V start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT studies usually implicitly assume that Nturbsubscript𝑁𝑡𝑢𝑟𝑏N_{turb}italic_N start_POSTSUBSCRIPT italic_t italic_u italic_r italic_b end_POSTSUBSCRIPT is constant or does not vary too much within an individual cloud (e.g., Ossenkopf & Mac Low, 2002; Liu et al., 2023b). Considering that each line tracer mainly traces a layer of gas with densities between ncsubscript𝑛𝑐n_{c}italic_n start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT and 100ncsubscript𝑛𝑐n_{c}italic_n start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, the validity of this assumption could be unclear.

Figure 9(b) shows the velocity centroid dispersion-radius relation. The σcrsubscript𝜎𝑐𝑟\sigma_{c}-ritalic_σ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT - italic_r relation in each region is complex and we refrain from interpreting them in detail in this work. The σcrsubscript𝜎𝑐𝑟\sigma_{c}-ritalic_σ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT - italic_r relations from different instruments and lines are not continuous as well. The σcsubscript𝜎𝑐\sigma_{c}italic_σ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT in most clumps traced by JCMT is comparable to σcsubscript𝜎𝑐\sigma_{c}italic_σ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT in the environmental gas traced by FCRAO-14m, but lower than the σcsubscript𝜎𝑐\sigma_{c}italic_σ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT of the dense cores traced by ALMA. If Nturb=L/λcsubscript𝑁𝑡𝑢𝑟𝑏𝐿subscript𝜆𝑐N_{turb}=\sqrt{L/\lambda_{c}}italic_N start_POSTSUBSCRIPT italic_t italic_u italic_r italic_b end_POSTSUBSCRIPT = square-root start_ARG italic_L / italic_λ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG is similar for different line data or Nturbsubscript𝑁𝑡𝑢𝑟𝑏N_{turb}italic_N start_POSTSUBSCRIPT italic_t italic_u italic_r italic_b end_POSTSUBSCRIPT is larger for higher-resolution data, the observed σcrsubscript𝜎𝑐𝑟\sigma_{c}-ritalic_σ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT - italic_r relation could suggest that the early-stage massive star formation activities have already increased the non-thermal motions in small-scale and high-density regions in IRDC G28.34. However, we cannot rule out the possibility that the velocity centroid dispersion at a smaller r𝑟ritalic_r is less averaged due to a smaller Nturbsubscript𝑁𝑡𝑢𝑟𝑏N_{turb}italic_N start_POSTSUBSCRIPT italic_t italic_u italic_r italic_b end_POSTSUBSCRIPT.

In summary, we find the velocity statistics are not continuous at different scales revealed by different instruments and lines, nor universal in different regions. However, due to the unsolved issues on the LOS length of linewidth and the averaging of velocity centroids, we cannot rule out the possibility that the actual multi-scale velocity statistics in G28.34 could still follow a continuous and universal power-law relation. Solving those issues requires the application of more advanced analysis methods (e.g., Lazarian & Pogosyan, 2006).

3.4 Magnetic field strength

The Davis-Chandrasekhar-Fermi (DCF) method (Davis, 1951; Chandrasekhar & Fermi, 1953) and its modified forms have been widely used to estimate the plane-of-sky (POS) magnetic field strength in molecular clouds. The basic assumptions of the DCF method are: there is an equipartition between the transverse (i.e., perpendicular to the ordered field) turbulent magnetic and kinetic energies; the turbulence is isotropic; and the turbulent-to-ordered or turbulent-to-total magnetic field ratio can be traced with the statistics of magnetic field orientations. Under these assumptions, the POS ordered and total magnetic field strengths are given by

B0=μ0ρσvBt/B0subscript𝐵0subscript𝜇0𝜌subscript𝜎𝑣subscript𝐵tsubscript𝐵0B_{0}=\sqrt{\mu_{0}\rho}\frac{\sigma_{v}}{B_{\mathrm{t}}/B_{0}}italic_B start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = square-root start_ARG italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ρ end_ARG divide start_ARG italic_σ start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT end_ARG start_ARG italic_B start_POSTSUBSCRIPT roman_t end_POSTSUBSCRIPT / italic_B start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG (11)

and

B=μ0ρσvBt/B,𝐵subscript𝜇0𝜌subscript𝜎𝑣subscript𝐵t𝐵B=\sqrt{\mu_{0}\rho}\frac{\sigma_{v}}{B_{\mathrm{t}}/B},italic_B = square-root start_ARG italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ρ end_ARG divide start_ARG italic_σ start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT end_ARG start_ARG italic_B start_POSTSUBSCRIPT roman_t end_POSTSUBSCRIPT / italic_B end_ARG , (12)

respectively, where μ0subscript𝜇0\mu_{0}italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the permeability of vacuum, ρ=μH2mHnH2𝜌subscript𝜇subscriptH2subscript𝑚Hsubscript𝑛subscriptH2\rho=\mu_{\mathrm{H_{2}}}m_{\mathrm{H}}n_{\mathrm{H_{2}}}italic_ρ = italic_μ start_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_m start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT is the gas density, nH2subscript𝑛subscriptH2n_{\mathrm{H_{2}}}italic_n start_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT is the volume density, and σvsubscript𝜎v\sigma_{\mathrm{v}}italic_σ start_POSTSUBSCRIPT roman_v end_POSTSUBSCRIPT is the line-of-sight turbulent velocity dispersion. In small angle approximation (i.e., the ordered magnetic field is prominent), the POS turbulent-to-ordered field ratio (Bt/B0subscript𝐵tsubscript𝐵0B_{\mathrm{t}}/B_{0}italic_B start_POSTSUBSCRIPT roman_t end_POSTSUBSCRIPT / italic_B start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT) or turbulent-to-total field ratio (Bt/Bsubscript𝐵t𝐵B_{\mathrm{t}}/Bitalic_B start_POSTSUBSCRIPT roman_t end_POSTSUBSCRIPT / italic_B) are usually estimated with the angular dispersion (σθsubscript𝜎𝜃\sigma_{\theta}italic_σ start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT) of POS magnetic field orientations: Bt/B0Bt/Bσθsimilar-tosubscript𝐵tsubscript𝐵0subscript𝐵t𝐵similar-tosubscript𝜎𝜃B_{\mathrm{t}}/B_{0}\sim B_{\mathrm{t}}/B\sim\sigma_{\theta}italic_B start_POSTSUBSCRIPT roman_t end_POSTSUBSCRIPT / italic_B start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∼ italic_B start_POSTSUBSCRIPT roman_t end_POSTSUBSCRIPT / italic_B ∼ italic_σ start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT. A recent review of the DCF method can be found in Liu et al. (2022a, b).

3.4.1 Environmental gas

We estimate the magnetic field strength in the environmental gas of IRDC G28.34 with the Planck polarization data and the FCRAO-14m 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO (1-0) data. As demonstrated later (see Section 3.5.2), the Planck polarization detection toward G28.34 mainly traces the emission from the surrounding gas of this IRDC. Thus it is appropriate to use the Planck polarization map to obtain the magnetic field information of the cloud environment. The radius of the polarization detection region must be greater-than-or-equivalent-to\gtrsim2 of the beam size to obtain meaningful statistics of the turbulent magnetic field (Liu et al., 2021). Thus we consider a circular region with a radius of r=𝑟absentr=italic_r =15 pc centered at G28.34. The angular dispersion of the Planck magnetic field is σθ=3.9°subscript𝜎𝜃3.9°\sigma_{\theta}=3.9\arcdegitalic_σ start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT = 3.9 ° within the considered region. The almost straight POS magnetic field lines at 15-pc scale and the small angular dispersion imply B0Bsimilar-tosubscript𝐵0𝐵B_{0}\sim Bitalic_B start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∼ italic_B. As mentioned in Section 3.2.3, the density estimated from the 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO (1-0) data tends to be underestimated while the density estimated from the Planck map tends to be overestimated. Thus we adopt the density extrapolated from the power-law relation (n=6398𝑛6398n=6398italic_n = 6398 cm(r/pc)1.413{}^{-3}(r/\mathrm{pc})^{-1.41}start_FLOATSUPERSCRIPT - 3 end_FLOATSUPERSCRIPT ( italic_r / roman_pc ) start_POSTSUPERSCRIPT - 1.41 end_POSTSUPERSCRIPT) fitted from the COMB data and ALMA data. At r=15𝑟15r=15italic_r = 15 pc, we obtain nH2144similar-tosubscript𝑛subscriptH2144n_{\mathrm{H_{2}}}\sim 144italic_n start_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT ∼ 144 cm33{}^{-3}start_FLOATSUPERSCRIPT - 3 end_FLOATSUPERSCRIPT. Similar to the turbulent kinetic field, the turbulent magnetic field as well as its ratio with respect to the ordered or total field could be underestimated by a factor of NBsubscript𝑁𝐵\sqrt{N_{B}}square-root start_ARG italic_N start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT end_ARG if there are NBsubscript𝑁𝐵N_{B}italic_N start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT independent magnetic turbulent cells along the LOS (Zweibel, 1990). The 3D unaveraged turbulent magnetic and kinetic fields within the considered region are not directly measurable from observations. As the DCF method assumes equipartition between turbulent magnetic and kinetic energies, it is reasonable to further assume that the turbulent magnetic and kinetic fields are averaged by a similar extent (i.e., NBNturbsimilar-tosubscript𝑁𝐵subscript𝑁𝑡𝑢𝑟𝑏N_{B}\sim N_{turb}italic_N start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ∼ italic_N start_POSTSUBSCRIPT italic_t italic_u italic_r italic_b end_POSTSUBSCRIPT) due to the LOS signal integration. Although Nturbsubscript𝑁𝑡𝑢𝑟𝑏N_{turb}italic_N start_POSTSUBSCRIPT italic_t italic_u italic_r italic_b end_POSTSUBSCRIPT and NBsubscript𝑁𝐵N_{B}italic_N start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT may not be accurately measured from our observational data, the corrections of the LOS signal integration effect for the turbulent kinetic field (as traced by σcsubscript𝜎𝑐\sigma_{c}italic_σ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT) and turbulent magnetic field (as traced by σθsubscript𝜎𝜃\sigma_{\theta}italic_σ start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT) could cancel out with (Nturbσc)/(NBσθ)=σc/σθsubscript𝑁𝑡𝑢𝑟𝑏subscript𝜎𝑐subscript𝑁𝐵subscript𝜎𝜃subscript𝜎𝑐subscript𝜎𝜃(\sqrt{N_{turb}}\sigma_{c})/(\sqrt{N_{B}}\sigma_{\theta})=\sigma_{c}/\sigma_{\theta}( square-root start_ARG italic_N start_POSTSUBSCRIPT italic_t italic_u italic_r italic_b end_POSTSUBSCRIPT end_ARG italic_σ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ) / ( square-root start_ARG italic_N start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT end_ARG italic_σ start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT ) = italic_σ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_σ start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT (Cho & Yoo, 2016). Within our studied region, the velocity centroid dispersion is σc=0.54subscript𝜎𝑐0.54\sigma_{c}=0.54italic_σ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 0.54 km s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT. Incorporating the correction for the LOS signal integration, the magnetic field strength of the environmental gas within the 15-pc region is given by BB0μ0ρσc/σθ0.074similar-to𝐵subscript𝐵0similar-tosubscript𝜇0𝜌subscript𝜎𝑐subscript𝜎𝜃similar-to0.074B\sim B_{0}\sim\sqrt{\mu_{0}\rho}\sigma_{c}/\sigma_{\theta}\sim 0.074italic_B ∼ italic_B start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∼ square-root start_ARG italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ρ end_ARG italic_σ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_σ start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT ∼ 0.074 mG. Alternatively, using a model with nH2100similar-tosubscript𝑛subscriptH2100n_{\mathrm{H_{2}}}\sim 100italic_n start_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT ∼ 100 cm33{}^{-3}start_FLOATSUPERSCRIPT - 3 end_FLOATSUPERSCRIPT and l=8𝑙8l=8italic_l = 8 pc, Ostriker et al. (2001) numerically investigated the uncertainties in the DCF estimation of clouds and derived a correction factor of 0.5 (hereafter the Ost01 correction factor). As the density of their simulation is comparable to that of the region of our interest, we apply the Ost01 correction factor for an alternative magnetic field strength estimation. Adopting σv=3.93subscript𝜎𝑣3.93\sigma_{v}=3.93italic_σ start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT = 3.93 km s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT at r=15𝑟15r=15italic_r = 15 pc, we obtain BB00.5μ0ρσv/σθ0.27similar-to𝐵subscript𝐵0similar-to0.5subscript𝜇0𝜌subscript𝜎𝑣subscript𝜎𝜃similar-to0.27B\sim B_{0}\sim 0.5\sqrt{\mu_{0}\rho}\sigma_{v}/\sigma_{\theta}\sim 0.27italic_B ∼ italic_B start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∼ 0.5 square-root start_ARG italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ρ end_ARG italic_σ start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT / italic_σ start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT ∼ 0.27 mG. Note that the critical density of 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO (1-0) is an order of magnitude higher than the density of our considered region, so the turbulent velocity dispersion here may be underestimated, which might lead to underestimation of the magnetic field strength as well.

At the considered scale, additional uncertainty on the estimated field strength could come from the anisotropy of MHD turbulence when gravity is negligible (Lazarian et al., 2022). The correction for the turbulence anisotropy requires knowledge of the mean-field inclination and the fraction of turbulence modes. However, both parameters cannot be easily measured with our existing data. Thus we are unable to correct this effect in this work.

Some recent theoretical and numerical studies suggest that in a non-gravitational and very sub-Alfvénic environment, the turbulent kinetic energy is in equipartition with the coupling-term magnetic field energy fluctuation rather than with the turbulent magnetic energy (Federrath, 2016; Skalidis & Tassis, 2021; Beattie et al., 2022), which could bring another uncertainty to the DCF method. However, it has been debated whether their proposed coupling-term energy equipartition is valid for molecular clouds (Lazarian et al., 2022; Li et al., 2022; Liu et al., 2022a, b). Firstly, the physical interpretation for including the coupling-term field in the energy equipartition is unclear. Even if the reference velocity is set as the cloud average velocity, the substructures moving at a different velocity from the cloud average velocity or the ordered non-turbulent motions (e.g., infall or rotation) due to star formation activities could still generate a significant amount of coupling-term velocity field fluctuation within molecular clouds. There is no reason to only consider the coupling-term magnetic field but not consider the coupling-term velocity field. Secondly, all the numerical works supporting the coupling-term energy equipartition have adopted the whole simulation-averaged mean magnetic field, which represents the interstellar medium (ISM) magnetic field at approximately the turbulence injection scale (similar-to\sim100 pc), in the calculation of the coupling-term magnetic field. While it is probably fine to adopt the global mean field as the local mean field in an extremely sub-Alfvénic case (e.g., A=0.01subscript𝐴0.01\mathcal{M}_{A}=0.01caligraphic_M start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT = 0.01, Beattie et al., 2022), the star-forming 10-pc clouds are only trans-to-sub-Alfvénic (e.g., A0.6greater-than-or-equivalent-tosubscript𝐴0.6\mathcal{M}_{A}\gtrsim 0.6caligraphic_M start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ≳ 0.6, Hu et al., 2019). In such cases, the 10-pc local mean field could have significantly deviated from the 100-pc global mean field and thus it would be inappropriate to adopt the global mean field of the ISM in the calculation of the cloud energetics. In summary, more work needs to be done to understand the validity of the coupling-term energy equipartition assumption in molecular clouds.

3.4.2 Clump

For the dense clumps revealed by JCMT observations, the magnetic field shows non-linear ordered field structures due to gravitational effects and star formation activities, which overestimates the angular dispersion that should be only attributed to turbulence. In G28.34, all clumps have σθ>25°subscript𝜎𝜃25°\sigma_{\theta}>25\arcdegitalic_σ start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT > 25 ° within r=1𝑟1r=1italic_r = 1 pc, but using σθsubscript𝜎𝜃\sigma_{\theta}italic_σ start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT to estimate B𝐵Bitalic_B is only valid when σθ<25°subscript𝜎𝜃25°\sigma_{\theta}<25\arcdegitalic_σ start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT < 25 ° in both low and high density regions (Ostriker et al., 2001; Liu et al., 2021). To account for the ordered field contribution, we use the angular dispersion function (ADF) method (Hildebrand et al., 2009; Houde et al., 2009, 2016), a modified DCF method, to estimate the magnetic field strength of these clumps.

The ADF accounting for the ordered field contribution, beam-smoothing effect, and turbulent correlation effect is given by

1cos[ΔΦ(l)]Bt2B2×[1el2/2(lδ2+2lW2)]+a2l2,similar-to-or-equals1delimited-⟨⟩ΔΦ𝑙delimited-⟨⟩superscriptsubscript𝐵t2delimited-⟨⟩superscript𝐵2delimited-[]1superscript𝑒superscript𝑙22superscriptsubscript𝑙𝛿22superscriptsubscript𝑙𝑊2superscriptsubscript𝑎2superscript𝑙21-\langle\cos[\Delta\Phi(l)]\rangle\simeq\frac{\langle B_{\mathrm{t}}^{2}% \rangle}{\langle B^{2}\rangle}\times[1-e^{-l^{2}/2(l_{\delta}^{2}+2l_{W}^{2})}% ]+a_{2}^{\prime}l^{2},1 - ⟨ roman_cos [ roman_Δ roman_Φ ( italic_l ) ] ⟩ ≃ divide start_ARG ⟨ italic_B start_POSTSUBSCRIPT roman_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ end_ARG start_ARG ⟨ italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ end_ARG × [ 1 - italic_e start_POSTSUPERSCRIPT - italic_l start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 ( italic_l start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 2 italic_l start_POSTSUBSCRIPT italic_W end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_POSTSUPERSCRIPT ] + italic_a start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_l start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (13)

where ΔΦ(l)ΔΦ𝑙\Delta\Phi(l)roman_Δ roman_Φ ( italic_l ) is the angular difference of two position angles separated by l𝑙litalic_l, lW=lbeam/8ln2subscript𝑙𝑊subscript𝑙beam82l_{W}=l_{\mathrm{beam}}/\sqrt{8\ln{2}}italic_l start_POSTSUBSCRIPT italic_W end_POSTSUBSCRIPT = italic_l start_POSTSUBSCRIPT roman_beam end_POSTSUBSCRIPT / square-root start_ARG 8 roman_ln 2 end_ARG is the standard deviation of the beam size, a2l2superscriptsubscript𝑎2superscript𝑙2a_{2}^{\prime}l^{2}italic_a start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_l start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT is the second-order term of the Taylor expansion for the ordered field, and lδsubscript𝑙𝛿l_{\delta}italic_l start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT is the turbulent correlation length. Here the POS turbulent-to-total magnetic energy ratio Bt2/B2delimited-⟨⟩superscriptsubscript𝐵t2delimited-⟨⟩superscript𝐵2\langle B_{\mathrm{t}}^{2}\rangle/\langle B^{2}\rangle⟨ italic_B start_POSTSUBSCRIPT roman_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ / ⟨ italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ does not consider the effect of LOS signal integration. Note that the second-order term a2l2superscriptsubscript𝑎2superscript𝑙2a_{2}^{\prime}l^{2}italic_a start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_l start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT is only valid to represent the ordered field at small l𝑙litalic_l when the higher-order terms are negligible. Also note that polarization observations can only trace the magnetic field orientation with a 180°°\arcdeg° ambiguity, so ΔΦ(l)ΔΦ𝑙\Delta\Phi(l)roman_Δ roman_Φ ( italic_l ) is constrained to be within [90,90]9090[-90,90][ - 90 , 90 ] degrees (Hildebrand et al., 2009; Houde et al., 2009). Because the actual magnetic field direction angle in the range of -180°°\arcdeg° to 180°°\arcdeg° is approximated by the position angle in the range of -90°°\arcdeg° to 90°°\arcdeg°, the ADF method implicitly assumes that the turbulent field is smaller than the ordered field (i.e., sub-Alfvénic).

Figure 10: Angular dispersion functions for MM4, MM6, MM9, and MM10. The diamond symbols represent the observed data points. The error bars indicate the statistical uncertainties propagated from the observational uncertainty (Houde et al., 2009). The blue dashed line indicates the best-fitted results. The cyan dashed line shows the large-scale component (Bt2/B2+a2l2delimited-⟨⟩superscriptsubscript𝐵t2delimited-⟨⟩superscript𝐵2superscriptsubscript𝑎2superscript𝑙2\langle B_{\mathrm{t}}^{2}\rangle/\langle B^{2}\rangle+a_{2}^{\prime}l^{2}⟨ italic_B start_POSTSUBSCRIPT roman_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ / ⟨ italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ + italic_a start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_l start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT) of the best fit. The horizontal dashed line indicates the ADF value corresponding to a random field (0.36, Liu et al., 2021).

Figure 10 shows the ADFs for clumps MM4, MM6, MM9, and MM10. We adopt a binning interval of lbin=lbeam/2subscript𝑙binsubscript𝑙beam2l_{\mathrm{bin}}=l_{\mathrm{beam}}/2italic_l start_POSTSUBSCRIPT roman_bin end_POSTSUBSCRIPT = italic_l start_POSTSUBSCRIPT roman_beam end_POSTSUBSCRIPT / 2 for the distance lag. We fit each ADF via reduced χ2superscript𝜒2\chi^{2}italic_χ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT minimization. We fit the ADFs over different maximum l𝑙litalic_l ranges and obtain the best fit with the smallest reduced χ2superscript𝜒2\chi^{2}italic_χ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. The best fitted results are Bt2/B20.5=0.25,0.26,0.31,0.22delimited-⟨⟩superscriptsubscript𝐵t2superscriptdelimited-⟨⟩superscript𝐵20.50.250.260.310.22\langle B_{\mathrm{t}}^{2}\rangle/\langle B^{2}\rangle^{0.5}=0.25,0.26,0.31,0.22⟨ italic_B start_POSTSUBSCRIPT roman_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ / ⟨ italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUPERSCRIPT 0.5 end_POSTSUPERSCRIPT = 0.25 , 0.26 , 0.31 , 0.22 for MM4, MM6, MM9, and MM10, respectively.

Due to the effect of LOS signal integration, the statistics of POS polarization position angles could underestimate the turbulent magnetic field. The ADF method proposes that the turbulent magnetic energy is averaged by a factor of

NB=(lδ2+2lW2)lΔ2πlδ3subscript𝑁𝐵superscriptsubscript𝑙𝛿22superscriptsubscript𝑙𝑊2subscript𝑙Δ2𝜋superscriptsubscript𝑙𝛿3N_{B}=\frac{(l_{\delta}^{2}+2l_{W}^{2})l_{\Delta}}{\sqrt{2\pi}l_{\delta}^{3}}italic_N start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT = divide start_ARG ( italic_l start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 2 italic_l start_POSTSUBSCRIPT italic_W end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) italic_l start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT end_ARG start_ARG square-root start_ARG 2 italic_π end_ARG italic_l start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG (14)

for single-dish data due to the LOS integration (Houde et al., 2009), where lΔsubscript𝑙Δl_{\Delta}italic_l start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT is the LOS cloud effective depth. Houde et al. (2009) suggested that lΔsubscript𝑙Δl_{\Delta}italic_l start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT could be estimated as the width at half of the maximum of the normalized autocorrelation function for the integrated normalized polarized flux. However, the normalized autocorrelation function of the integrated normalized polarized flux is greater than half of the maximum for all the clumps, so lΔsubscript𝑙Δl_{\Delta}italic_l start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT in our case cannot be derived in this way. Moreover, the numerical study by Liu et al. (2021) found that the ADF method may not work well for the effect of LOS signal integration. So we refrain from adopting the correction for this effect suggested by Houde et al. (2009).

Alternatively, we adopt the analytical corrections suggested by Cho & Yoo (2016) (CY16) or numerical corrections by Liu et al. (2021) (Liu21). Cho & Yoo (2016) suggested using the turbulent velocity centroid dispersion instead of the turbulent velocity dispersion in the DCF equation to account for the LOS averaging effect. Here we derive σcsubscript𝜎𝑐\sigma_{c}italic_σ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT from the 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO (3-2) line data because its critical density is closer to the clump densities. The velocity centroid dispersions are σc=subscript𝜎𝑐absent\sigma_{c}=italic_σ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT =0.34, 0.30, 0.22, and 0.23 km s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT for MM4, MM6, MM9, and MM10, respectively. Adopting n=6.3×103𝑛6.3superscript103n=6.3\times 10^{3}italic_n = 6.3 × 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT, 5.8×1035.8superscript1035.8\times 10^{3}5.8 × 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT, 5.7×1035.7superscript1035.7\times 10^{3}5.7 × 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT, and 6.0×1036.0superscript1036.0\times 10^{3}6.0 × 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT cm33{}^{-3}start_FLOATSUPERSCRIPT - 3 end_FLOATSUPERSCRIPT, we obtain Bμ0ρσc(Bt2/B2)0.5similar-to𝐵subscript𝜇0𝜌subscript𝜎𝑐superscriptdelimited-⟨⟩superscriptsubscript𝐵t2delimited-⟨⟩superscript𝐵20.5similar-toabsentB\sim\sqrt{\mu_{0}\rho}\sigma_{c}(\langle B_{\mathrm{t}}^{2}\rangle/\langle B^% {2}\rangle)^{-0.5}\simitalic_B ∼ square-root start_ARG italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ρ end_ARG italic_σ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( ⟨ italic_B start_POSTSUBSCRIPT roman_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ / ⟨ italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ ) start_POSTSUPERSCRIPT - 0.5 end_POSTSUPERSCRIPT ∼0.081, 0.065, 0.039, and 0.060 mG for MM4, MM6, MM9, and MM10, respectively. Here we assume that the velocity centroid dispersion is dominated by turbulent motions rather than non-turbulent motions. The magnetic field strength could be overestimated if σcsubscript𝜎𝑐\sigma_{c}italic_σ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT is mostly non-turbulent. On the other hand, Liu et al. (2021) has numerically derived the average correction factors for the clump-to-core scale magnetic field strength estimated with the ADF method using the velocity dispersion. The velocity dispersions are σv=subscript𝜎𝑣absent\sigma_{v}=italic_σ start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT =2.32, 2.54, 1.91, and 2.70 km s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT for MM4, MM6, MM9, and MM10, respectively. Adopting a numerical correction factor of 0.21 (with a 45% uncertainty, Liu et al., 2021), we obtain B0.21μ0ρσv(Bt2/B2)0.5similar-to𝐵0.21subscript𝜇0𝜌subscript𝜎𝑣superscriptdelimited-⟨⟩superscriptsubscript𝐵t2delimited-⟨⟩superscript𝐵20.5similar-toabsentB\sim 0.21\sqrt{\mu_{0}\rho}\sigma_{v}(\langle B_{\mathrm{t}}^{2}\rangle/% \langle B^{2}\rangle)^{-0.5}\simitalic_B ∼ 0.21 square-root start_ARG italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ρ end_ARG italic_σ start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT ( ⟨ italic_B start_POSTSUBSCRIPT roman_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ / ⟨ italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ ) start_POSTSUPERSCRIPT - 0.5 end_POSTSUPERSCRIPT ∼0.120, 0.120, 0.075, and 0.154 mG for MM4, MM6, MM9, and MM10, respectively. The magnetic field strengths of the clumps estimated with either the CY16 or Liu21 corrections are comparable to the magnetic field strength of the environmental gas, which means the magnetic field does not significantly scale with density in the low-density gas. This behavior is consistent with previous magnetic field strength estimations in other regions (Pattle & Fissel, 2019; Liu et al., 2022a, b). Using the DCF methods to derive the ordered field strength could have large uncertainties in self-gravitating regions (Liu et al., 2021), so we only derive the total field strength for the clumps.

Within self-gravitating clumps and cores, the uncertainty from anisotropic turbulence on the DCF method is negligible (Liu et al., 2022b). Thus we do not consider the correction for this effect. Because there is no evidence supporting coupling-term energy equipartition in self-gravitating regions (Liu et al., 2022b), we keep our assumption of turbulent energy equipartition. Note that if the clumps are super-Alfvénic, the turbulent magnetic energy could be smaller than the turbulent kinetic energy, and our derived magnetic field strengths for the clumps could be overestimated.

3.4.3 Core

From our ALMA observations, only one core (MM4-core4) has sufficient polarization detections to derive the magnetic field strength. Liu et al. (2020) has estimated the ordered magnetic field strength of MM4-core4 using the ADF method. The kinetic information adopted in Liu et al. (2020) was from the EVLA NH33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT line data (Wang et al., 2012), which had a different resolution and filtering scale larger than the ALMA data. Here we recalculate the magnetic field strength with the CY16 and Liu21 corrections and with the velocity information from the ALMA N22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPTD+{}^{+}start_FLOATSUPERSCRIPT + end_FLOATSUPERSCRIPT line data.

In MM4-core4, the velocity dispersion and velocity centroid dispersion of N22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPTD+{}^{+}start_FLOATSUPERSCRIPT + end_FLOATSUPERSCRIPT are 0.61 and 0.37 km s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT, respectively. We adopt the radius (r=0.053𝑟0.053r=0.053italic_r = 0.053 pc), density (n=1.1×106𝑛1.1superscript106n=1.1\times 10^{6}italic_n = 1.1 × 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT cm33{}^{-3}start_FLOATSUPERSCRIPT - 3 end_FLOATSUPERSCRIPT), and turbulent-to-total magnetic field strength ratio without correction for LOS integration ((Bt2/B2)0.51similar-tosuperscriptdelimited-⟨⟩superscriptsubscript𝐵t2delimited-⟨⟩superscript𝐵20.51(\langle B_{\mathrm{t}}^{2}\rangle/\langle B^{2}\rangle)^{0.5}\sim 1( ⟨ italic_B start_POSTSUBSCRIPT roman_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ / ⟨ italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ ) start_POSTSUPERSCRIPT 0.5 end_POSTSUPERSCRIPT ∼ 1. i.e., the total field is dominated by the turbulent field) estimated in Liu et al. (2020). Adopting the CY16 and Liu21 corrections, we obtain Bμ0ρσc(Bt2/B2)0.5similar-to𝐵subscript𝜇0𝜌subscript𝜎𝑐superscriptdelimited-⟨⟩superscriptsubscript𝐵t2delimited-⟨⟩superscript𝐵20.5similar-toabsentB\sim\sqrt{\mu_{0}\rho}\sigma_{c}(\langle B_{\mathrm{t}}^{2}\rangle/\langle B^% {2}\rangle)^{-0.5}\simitalic_B ∼ square-root start_ARG italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ρ end_ARG italic_σ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( ⟨ italic_B start_POSTSUBSCRIPT roman_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ / ⟨ italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ ) start_POSTSUPERSCRIPT - 0.5 end_POSTSUPERSCRIPT ∼0.29 mG and B0.21μ0ρσv(Bt2/B2)0.5similar-to𝐵0.21subscript𝜇0𝜌subscript𝜎𝑣superscriptdelimited-⟨⟩superscriptsubscript𝐵t2delimited-⟨⟩superscript𝐵20.5similar-toabsentB\sim 0.21\sqrt{\mu_{0}\rho}\sigma_{v}(\langle B_{\mathrm{t}}^{2}\rangle/% \langle B^{2}\rangle)^{-0.5}\simitalic_B ∼ 0.21 square-root start_ARG italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ρ end_ARG italic_σ start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT ( ⟨ italic_B start_POSTSUBSCRIPT roman_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ / ⟨ italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ ) start_POSTSUPERSCRIPT - 0.5 end_POSTSUPERSCRIPT ∼0.10 mG, respectively. It should be noted that both the CY16 and Liu21 corrections may not well account for the LOS integration effect at <<<0.1 pc and n>106𝑛superscript106n>10^{6}italic_n > 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT cm33{}^{-3}start_FLOATSUPERSCRIPT - 3 end_FLOATSUPERSCRIPT (Liu et al., 2021). Also, note that the turbulent velocity motions may be overestimated due to the existence of non-turbulent non-thermal motions. And if the core is super-Alfvénic, the energy equipartition assumption could break down. Those above-mentioned factors could all overestimate the B strength for MM4-core4. On the other hand, the ALMA observations filter out the large-scale emission, but Liu et al. (2021) did not perform the filtering when deriving the velocity dispersion from the simulations. This inconsistency could underestimate the B strength estimated from the Liu21 correction when using the line velocity dispersion from interferometric observations, which could explain the smaller B value from the Liu21 correction than that from the Cy16 correction. Overall, the B value in core MM4-core4 is comparable to or only slightly larger than that in clump MM4, which may suggest that gravity does not significantly compress and amplify the magnetic field strength in the clumps in the early massive star formation stage.

3.5 Velocity Gradient Technique

3.5.1 Alfvénic Mach number

Using the angular dispersion of polarization position angles to derive the Alfvénic Mach number Asubscript𝐴\mathcal{M}_{A}caligraphic_M start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT requires corrections for the LOS signal integration and other effects. It is inappropriate to just equal the angular dispersion and Asubscript𝐴\mathcal{M}_{A}caligraphic_M start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT as some previous studies have done. Based on the property of MHD turbulence (Goldreich & Sridhar, 1995), Lazarian et al. (2018) proposed an alternative method (Velocity Gradient Technique, VGT) to estimate Asubscript𝐴\mathcal{M}_{A}caligraphic_M start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT in non-self-gravitating regions with the statistics of velocity gradient orientation (θVGsubscript𝜃VG\theta_{\mathrm{VG}}italic_θ start_POSTSUBSCRIPT roman_VG end_POSTSUBSCRIPT) from molecular line observations. The advantage of VGT is that the velocity gradient orientation histogram is independent of the LOS integration (Lazarian et al., 2018). It was numerically shown that the top-to-bottom ratio (Ntop/Nbotsubscript𝑁𝑡𝑜𝑝subscript𝑁𝑏𝑜𝑡N_{top}/N_{bot}italic_N start_POSTSUBSCRIPT italic_t italic_o italic_p end_POSTSUBSCRIPT / italic_N start_POSTSUBSCRIPT italic_b italic_o italic_t end_POSTSUBSCRIPT) of the velocity gradient orientation histogram has the relation with Asubscript𝐴\mathcal{M}_{A}caligraphic_M start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT in a sub-Alfvénic region (Hu et al., 2021)

A1.6(Ntop/Nbot)10.6.similar-to-or-equalssubscript𝐴1.6superscriptsubscript𝑁𝑡𝑜𝑝subscript𝑁𝑏𝑜𝑡10.6\mathcal{M}_{A}\simeq 1.6(N_{top}/N_{bot})^{\frac{1}{-0.6}}.caligraphic_M start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ≃ 1.6 ( italic_N start_POSTSUBSCRIPT italic_t italic_o italic_p end_POSTSUBSCRIPT / italic_N start_POSTSUBSCRIPT italic_b italic_o italic_t end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG - 0.6 end_ARG end_POSTSUPERSCRIPT . (15)

The Top (Ntopsubscript𝑁𝑡𝑜𝑝N_{top}italic_N start_POSTSUBSCRIPT italic_t italic_o italic_p end_POSTSUBSCRIPT) and Bottom (Nbotsubscript𝑁𝑏𝑜𝑡N_{bot}italic_N start_POSTSUBSCRIPT italic_b italic_o italic_t end_POSTSUBSCRIPT) values of the histogram can be obtained by fitting a Gaussian profile (i.e., (NtopNbot)exp(α(θVGθ0)2)+Nbotsubscript𝑁𝑡𝑜𝑝subscript𝑁𝑏𝑜𝑡𝛼superscriptsubscript𝜃VGsubscript𝜃02subscript𝑁𝑏𝑜𝑡(N_{top}-N_{bot})\exp(-\alpha(\theta_{\mathrm{VG}}-\theta_{\mathrm{0}})^{2})+N% _{bot}( italic_N start_POSTSUBSCRIPT italic_t italic_o italic_p end_POSTSUBSCRIPT - italic_N start_POSTSUBSCRIPT italic_b italic_o italic_t end_POSTSUBSCRIPT ) roman_exp ( - italic_α ( italic_θ start_POSTSUBSCRIPT roman_VG end_POSTSUBSCRIPT - italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) + italic_N start_POSTSUBSCRIPT italic_b italic_o italic_t end_POSTSUBSCRIPT, where α𝛼\alphaitalic_α and θ0subscript𝜃0\theta_{\mathrm{0}}italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT are coefficients) for the Vcsubscript𝑉𝑐V_{c}italic_V start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT histogram.

Figure 11: Histogram of the centered velocity gradient orientation for the FCRAO-14m 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO (1-0) line data within the r=15𝑟15r=15italic_r = 15 pc region. The blue line indicates the best-fitted result.

Figure 11 shows the histogram of the centered velocity gradient orientation (θVGθ0subscript𝜃VGsubscript𝜃0\theta_{\mathrm{VG}}-\theta_{\mathrm{0}}italic_θ start_POSTSUBSCRIPT roman_VG end_POSTSUBSCRIPT - italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT) for the FCRAO-14m 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO (1-0) line data within the r=15𝑟15r=15italic_r = 15 pc region centered at G28.34. We fit the histogram with a Gaussian profile and obtain A=0.74subscript𝐴0.74\mathcal{M}_{A}=0.74caligraphic_M start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT = 0.74, which is slightly sub-Alfvénic.

Within molecular clouds, the velocity gradient could be significantly affected by gravitational effects and star formation activities, which makes the property of velocity gradient statistics overall deviate from what is expected for pure MHD turbulence. Thus we refrain from applying the VGT to smaller scales.

3.5.2 Magnetic field orientation

The Planck polarization map probes all the LOS structures, which makes it hard to separate the emission from the cloud environment and the foreground/background. Here we independently derive the magnetic field orientation from the FCRAO-14m 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO (1-0) data with the VGT to confirm whether the Planck polarization toward G28.34 mainly traces emission from the cloud environment. The basic principle behind VGT is that the magnetized turbulence eddy is anisotropic and elongated along the magnetic field, so the velocity gradient of MHD turbulence should be perpendicular to the magnetic field (Goldreich & Sridhar, 1995) in the absence of gravity. By selecting different velocity ranges, this technique can separate LOS velocity components corresponding to different distances, thus avoiding contamination from the foreground/background.

We implement the velocity channel gradients (VChG, Lazarian & Yuen, 2018) on the FCRAO-14m data with approaches similar to Hu et al. (2022) and Hu & Lazarian (2023): (1) We convolve the line intensity map at each velocity channel with a Sobel kernel to obtain the intensity gradient map of each channel (i.e., channel gradient). Pixels with S/N<<<5 are not considered in this step. (2) In the POS, we bin the Planck map over 3×\times×3 pixels and perform sub-block averaging (Yuen & Lazarian, 2017a) for the line channel gradients within the area of each binned Planck pixel. The ratio between the block size and the line spatial resolution is similar-to\sim6.7, which is similar to previous VGT studies (e.g., Hu et al., 2022) and is sufficient for the statistics required by sub-block averaging. With a circular Gaussian fit for the histogram of channel gradients, we find the averaged gradient orientation ΨgssubscriptΨ𝑔𝑠\Psi_{gs}roman_Ψ start_POSTSUBSCRIPT italic_g italic_s end_POSTSUBSCRIPT for each block. Note that due to sub-block averaging, the resolution of the spectral data is degraded to the block size (similar-to\sim5\arcmin). (3) Integrating channels along the LOS, we calculate pseudo-Stokes parameters Qgsubscript𝑄𝑔Q_{g}italic_Q start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT and Ugsubscript𝑈𝑔U_{g}italic_U start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT with

Qg=ΣiNchIicos(2Ψgs,i),subscript𝑄𝑔superscriptsubscriptΣ𝑖subscript𝑁chsubscript𝐼𝑖2subscriptΨ𝑔𝑠𝑖Q_{g}=\Sigma_{i}^{N_{\mathrm{ch}}}I_{i}\cos(2\Psi_{gs,i}),italic_Q start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT = roman_Σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT roman_ch end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_I start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT roman_cos ( 2 roman_Ψ start_POSTSUBSCRIPT italic_g italic_s , italic_i end_POSTSUBSCRIPT ) , (16)
Ug=ΣiNchIisin(2Ψgs,i),subscript𝑈𝑔superscriptsubscriptΣ𝑖subscript𝑁chsubscript𝐼𝑖2subscriptΨ𝑔𝑠𝑖U_{g}=\Sigma_{i}^{N_{\mathrm{ch}}}I_{i}\sin(2\Psi_{gs,i}),italic_U start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT = roman_Σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT roman_ch end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_I start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT roman_sin ( 2 roman_Ψ start_POSTSUBSCRIPT italic_g italic_s , italic_i end_POSTSUBSCRIPT ) , (17)

where Iisubscript𝐼𝑖I_{i}italic_I start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is the block-averaged intensity. Here we integrate the velocity channels within three different velocity ranges: the cloud velocity range (from 71 to 86 km s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT), ±plus-or-minus\pm±5 km s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT (from 74 to 84 km s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT), and ±plus-or-minus\pm±3 km s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT (from 76 to 82 km s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT). (4) The orientation of VChG is given by θVChG=0.5arctan(Ug/Qg)subscript𝜃VChG0.5subscript𝑈𝑔subscript𝑄𝑔\theta_{\mathrm{VChG}}=0.5\arctan(U_{g}/Q_{g})italic_θ start_POSTSUBSCRIPT roman_VChG end_POSTSUBSCRIPT = 0.5 roman_arctan ( italic_U start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT / italic_Q start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ). The pseudo-magnetic field orientation is then given by θB,VGT=θVChG+90°subscript𝜃BVGTsubscript𝜃VChG90°\theta_{\mathrm{B,VGT}}=\theta_{\mathrm{VChG}}+90\arcdegitalic_θ start_POSTSUBSCRIPT roman_B , roman_VGT end_POSTSUBSCRIPT = italic_θ start_POSTSUBSCRIPT roman_VChG end_POSTSUBSCRIPT + 90 °.

Figure 12: Pseudo magnetic field orientations inferred from VChG (red lines) within different velocity ranges (indicated in each panel) overlaid on the column density map. The Magnetic field orientations revealed by the Planck dust polarization are shown in black lines. The blue line indicates the average magnetic field orientation of the convolved JCMT polarization map. Line lengths are arbitrary. The white dashed circle marks the r=15𝑟15r=15italic_r = 15 pc region surrounding G28.34 within which we perform the DCF analysis. Within r=15𝑟15r=15italic_r = 15 pc, the pseudo magnetic field orientations derived from VGT within different velocity ranges are very similar.

Figure 12 compares the magnetic field orientations inferred from the VChG with those traced by the Planck dust polarization. The two orientations are well aligned with each other in bright regions near G28.34. As we only select the cloud velocity range in the VGT analysis, the field orientation inferred from VChG should not contain contributions from the foreground/background. Moreover, we convolve the JCMT maps to the same resolution as Planck, then estimate the average polarization position angle within 3\arcmin of the peak of the convolved total intensity map. The magnetic field orientation measured from the convolved JCMT map is only 9°°\arcdeg° different from the Planck measurement and 7°°\arcdeg° different from the VGT measurement at the nearest pixel (see Figure 12). The close alignment between measurements from VChG, JCMT, and Planck tends to suggest that the Planck dust polarization toward G28.34 mainly traces the emission from G28.34 and its close environment, but not the foreground/background. Yuen & Lazarian (2017b) suggested that the magnetic field orientation and the block-averaged velocity gradient could transit to a perpendicular alignment (i.e., re-rotation) when gravity becomes important in high-density regions. Although there are some slight angular differences between the field orientations revealed by VChG and Planck, overall there is a lack of evidence for perpendicular alignments toward the brightest pixels. This suggests that gravitational motions do not dominate in the large-scale diffuse gas. On the other hand, the field orientations traced by VChG and Planck show larger differences in weak-emission regions, which may suggest that the Planck dust polarization mainly traces the materials outside of the cloud velocity range in those regions or those weak-emission areas are significantly influenced by noise.

In Section 3.4.1, we used the angular dispersion measured from the Planck polarization data to estimate the magnetic field strength of the cloud environment. Alternatively, if we use the angular dispersion from the VGT measurements in the DCF calculations, the estimated magnetic field strength will decrease by a factor of similar-to\sim3. However, this alternative field strength estimation should be regarded as a lower limit. This is because the angular dispersion from VGT is expected to be affected by ordered velocity fields. i.e., although not dominant, the gravitational motions may have already contributed to the orientations of VChG, which increases the angular dispersion of VGT measurements.

3.6 Relative orientation analysis

Here we investigate the multi-scale physical properties of G28.34 with a synergistic local relative orientation analysis (Liu et al., 2023a) combining the approaches of the polarization-intensity gradient (Koch-Tang-Ho or KTH) method (Koch et al., 2012a) and the Histogram of Relative Orientation (HRO) analysis (Soler et al., 2013). Basically, we characterize the relative angle between magnetic fields and other orientations (column density gradient and local gravity) with the alignment measure (AM) parameter (Lazarian & Yuen, 2018). The AM𝐴𝑀AMitalic_A italic_M is given by

AM=cos(2ϕo1o2),𝐴𝑀delimited-⟨⟩2superscriptsubscriptitalic-ϕ𝑜1𝑜2AM=\langle\cos(2\phi_{o1}^{o2})\rangle,italic_A italic_M = ⟨ roman_cos ( 2 italic_ϕ start_POSTSUBSCRIPT italic_o 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_o 2 end_POSTSUPERSCRIPT ) ⟩ , (18)

where ϕo1o2=|θo1θo2|superscriptsubscriptitalic-ϕo1o2subscript𝜃o1subscript𝜃o2\phi_{\mathrm{o1}}^{\mathrm{o2}}=|\theta_{\mathrm{o1}}-\theta_{\mathrm{o2}}|italic_ϕ start_POSTSUBSCRIPT o1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT o2 end_POSTSUPERSCRIPT = | italic_θ start_POSTSUBSCRIPT o1 end_POSTSUBSCRIPT - italic_θ start_POSTSUBSCRIPT o2 end_POSTSUBSCRIPT | is the angle between two orientations and is in the range of 0 to 90°°\arcdeg°. The uncertainty of ϕo1o2superscriptsubscriptitalic-ϕo1o2\phi_{\mathrm{o1}}^{\mathrm{o2}}italic_ϕ start_POSTSUBSCRIPT o1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT o2 end_POSTSUPERSCRIPT is given by δϕo1o2=δθo12+δθo22𝛿superscriptsubscriptitalic-ϕo1o2𝛿superscriptsubscript𝜃o12𝛿superscriptsubscript𝜃o22\delta\phi_{\mathrm{o1}}^{\mathrm{o2}}=\sqrt{\delta\theta_{\mathrm{o1}}^{2}+% \delta\theta_{\mathrm{o2}}^{2}}italic_δ italic_ϕ start_POSTSUBSCRIPT o1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT o2 end_POSTSUPERSCRIPT = square-root start_ARG italic_δ italic_θ start_POSTSUBSCRIPT o1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_δ italic_θ start_POSTSUBSCRIPT o2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG. Data points with δϕ>10°𝛿italic-ϕ10°\delta\phi>10\arcdegitalic_δ italic_ϕ > 10 ° are excluded from our analysis. The uncertainty of AM𝐴𝑀AMitalic_A italic_M is given by (Liu et al., 2023a)

δAM=((cos(2ϕo1o2))2AM2+Σin(2sin(2ϕi)δϕi)2)/n,𝛿𝐴𝑀delimited-⟨⟩superscript2superscriptsubscriptitalic-ϕ𝑜1𝑜22𝐴superscript𝑀2superscriptsubscriptΣ𝑖superscript𝑛superscript22subscriptitalic-ϕ𝑖𝛿subscriptitalic-ϕ𝑖2superscript𝑛\delta AM=\sqrt{(\langle(\cos(2\phi_{o1}^{o2}))^{2}\rangle-AM^{2}+\Sigma_{i}^{% n^{\prime}}(2\sin(2\phi_{i})\delta\phi_{i})^{2})/n^{\prime}},italic_δ italic_A italic_M = square-root start_ARG ( ⟨ ( roman_cos ( 2 italic_ϕ start_POSTSUBSCRIPT italic_o 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_o 2 end_POSTSUPERSCRIPT ) ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ - italic_A italic_M start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + roman_Σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT ( 2 roman_sin ( 2 italic_ϕ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) italic_δ italic_ϕ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) / italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_ARG , (19)

where nsuperscript𝑛n^{\prime}italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT is the number of data points considered. For the Planck data, we consider area within r=15𝑟15r=15italic_r = 15 pc centered at G28.34. For the JCMT data, we consider area with S/N(I𝐼Iitalic_I)>>>5 and S/N(PI𝑃𝐼PIitalic_P italic_I)>>>2. For the ALMA data, we consider area with S/N(I𝐼Iitalic_I)>>>3 and S/N(PI𝑃𝐼PIitalic_P italic_I)>>>2. We calculate AM in 6, 10, and 15 different column density bins for the Planck, JCMT, and ALMA data, respectively. The typical numbers of pixels are similar-to\sim20, similar-to\sim20-80, and similar-to\sim30-180 per bin for the Planck, JCMT, and ALMA data, respectively.

3.6.1 Magnetic field versus column density gradient

Figure 13 shows the AM-N relation for the angle between the magnetic field (θBsubscript𝜃B\theta_{\mathrm{B}}italic_θ start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT) and column density gradient (θNGsubscript𝜃NG\theta_{\mathrm{NG}}italic_θ start_POSTSUBSCRIPT roman_NG end_POSTSUBSCRIPT). The column density gradient is obtained by applying a 3×3333\times 33 × 3 Sobel kernel to the column density map. The uncertainty of θNGsubscript𝜃NG\theta_{\mathrm{NG}}italic_θ start_POSTSUBSCRIPT roman_NG end_POSTSUBSCRIPT is derived following Liu et al. (2023a). As θNGsubscript𝜃NG\theta_{\mathrm{NG}}italic_θ start_POSTSUBSCRIPT roman_NG end_POSTSUBSCRIPT is perpendicular to the column density contour, studying the AM-N relation for ϕBNGsuperscriptsubscriptitalic-ϕ𝐵𝑁𝐺\phi_{B}^{NG}italic_ϕ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N italic_G end_POSTSUPERSCRIPT is equivalent to the HRO analysis (Soler et al., 2013). As seen in Figure 13, the relative orientation changes from a statistically slightly more perpendicular alignment (AMBNG0less-than-or-similar-to𝐴superscriptsubscript𝑀𝐵𝑁𝐺0AM_{B}^{NG}\lesssim 0italic_A italic_M start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N italic_G end_POSTSUPERSCRIPT ≲ 0) in the environment at low column densities as revealed by Planck to a slightly more parallel alignment (AMBNG0greater-than-or-equivalent-to𝐴superscriptsubscript𝑀𝐵𝑁𝐺0AM_{B}^{NG}\gtrsim 0italic_A italic_M start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N italic_G end_POSTSUPERSCRIPT ≳ 0) in G28.34 at intermediate column densities as revealed by JCMT. This transition can only be reproduced in trans-to-sub-Alfvénic simulations in numerical HRO studies (see a review of the HRO studies in Liu et al., 2022b), which suggests that G28.34 is in a trans-to-sub-Alfvénic environment. This is in agreement with the result of our VGT analysis in Section 3.5. The reasons for the different alignment at different column densities and the transition of alignment are still under debate (Liu et al., 2022b). As discussed below (see Section 3.6.2), the transition to AMBNG0greater-than-or-equivalent-to𝐴superscriptsubscript𝑀𝐵𝑁𝐺0AM_{B}^{NG}\gtrsim 0italic_A italic_M start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N italic_G end_POSTSUPERSCRIPT ≳ 0 may be related to the distortion of gravity. At high column densities revealed by ALMA, the two angles tend to transit back to AMBNG0similar-to𝐴superscriptsubscript𝑀𝐵𝑁𝐺0AM_{B}^{NG}\sim 0italic_A italic_M start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N italic_G end_POSTSUPERSCRIPT ∼ 0 as N𝑁Nitalic_N increases, which may be due to the influence of early massive star formation activities (e.g., infall, rotation, outflows, accretion, and et al.). Overall, the alignment between θBsubscript𝜃B\theta_{\mathrm{B}}italic_θ start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT and θNGsubscript𝜃NG\theta_{\mathrm{NG}}italic_θ start_POSTSUBSCRIPT roman_NG end_POSTSUBSCRIPT in IRDC G28.34 at different column densities is very similar to what we have found in the evolved massive star formation region NGC 6334 (Liu et al., 2023a).

Figure 13: Relative orientations (characterized by AM𝐴𝑀AMitalic_A italic_M. See Equation 18) between magnetic field (θBsubscript𝜃B\theta_{\mathrm{B}}italic_θ start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT) and column density gradient (θNGsubscript𝜃NG\theta_{\mathrm{NG}}italic_θ start_POSTSUBSCRIPT roman_NG end_POSTSUBSCRIPT) as a function of column density for Planck (left), JCMT (middle), and ALMA (right) observations. Due to the filtering of large-scale emissions for JCMT and ALMA, the absolute column densities from different instruments are not comparable. AM>0𝐴𝑀0AM>0italic_A italic_M > 0 and AM<0𝐴𝑀0AM<0italic_A italic_M < 0 indicate a preferentially parallel and perpendicular alignment, respectively.

3.6.2 Magnetic field versus local gravity

Figure 14 shows the AM-N relation for the angle between the magnetic field (θBsubscript𝜃B\theta_{\mathrm{B}}italic_θ start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT) and local gravity (θLGsubscript𝜃LG\theta_{\mathrm{LG}}italic_θ start_POSTSUBSCRIPT roman_LG end_POSTSUBSCRIPT). Considering gas mass at S/N(I𝐼Iitalic_I)>>>3, the 2D local gravity direction (θLGsubscript𝜃LG\theta_{\mathrm{LG}}italic_θ start_POSTSUBSCRIPT roman_LG end_POSTSUBSCRIPT) at position 𝒓𝒋subscript𝒓𝒋\boldsymbol{r_{j}}bold_italic_r start_POSTSUBSCRIPT bold_italic_j end_POSTSUBSCRIPT is calculated with the standard formula of gravitation:

𝒈𝒋(𝒓)=Gi=1nmimj|𝒓𝒋𝒓𝒊|2𝒆𝒊𝒋,subscript𝒈𝒋𝒓𝐺superscriptsubscript𝑖1𝑛subscript𝑚𝑖subscript𝑚𝑗superscriptsubscript𝒓𝒋subscript𝒓𝒊2subscript𝒆𝒊𝒋\boldsymbol{g_{j}(r)}=G\sum\limits_{i=1}^{n}\frac{m_{i}m_{j}}{|\boldsymbol{r_{% j}}-\boldsymbol{r_{i}}|^{2}}\boldsymbol{e_{ij}},bold_italic_g start_POSTSUBSCRIPT bold_italic_j end_POSTSUBSCRIPT bold_( bold_italic_r bold_) = italic_G ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT divide start_ARG italic_m start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_m start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG start_ARG | bold_italic_r start_POSTSUBSCRIPT bold_italic_j end_POSTSUBSCRIPT - bold_italic_r start_POSTSUBSCRIPT bold_italic_i end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG bold_italic_e start_POSTSUBSCRIPT bold_italic_i bold_italic_j end_POSTSUBSCRIPT , (20)

where 𝒆𝒊𝒋subscript𝒆𝒊𝒋\boldsymbol{e_{ij}}bold_italic_e start_POSTSUBSCRIPT bold_italic_i bold_italic_j end_POSTSUBSCRIPT is the direction between position 𝒓𝒊subscript𝒓𝒊\boldsymbol{r_{i}}bold_italic_r start_POSTSUBSCRIPT bold_italic_i end_POSTSUBSCRIPT and 𝒓𝒋subscript𝒓𝒋\boldsymbol{r_{j}}bold_italic_r start_POSTSUBSCRIPT bold_italic_j end_POSTSUBSCRIPT, G𝐺Gitalic_G is the gravitational constant, mjsubscript𝑚𝑗m_{j}italic_m start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT is the mass at position 𝒓𝒋subscript𝒓𝒋\boldsymbol{r_{j}}bold_italic_r start_POSTSUBSCRIPT bold_italic_j end_POSTSUBSCRIPT, and θLGsubscript𝜃LG\theta_{\mathrm{LG}}italic_θ start_POSTSUBSCRIPT roman_LG end_POSTSUBSCRIPT is the direction of 𝒈𝒋(𝒓)subscript𝒈𝒋𝒓\boldsymbol{g_{j}(r)}bold_italic_g start_POSTSUBSCRIPT bold_italic_j end_POSTSUBSCRIPT bold_( bold_italic_r bold_). Overall, the AM-N trend for ϕBLGsuperscriptsubscriptitalic-ϕ𝐵𝐿𝐺\phi_{B}^{LG}italic_ϕ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_L italic_G end_POSTSUPERSCRIPT is similar to the trend for ϕBNGsuperscriptsubscriptitalic-ϕ𝐵𝑁𝐺\phi_{B}^{NG}italic_ϕ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N italic_G end_POSTSUPERSCRIPT, where AMBLG𝐴superscriptsubscript𝑀𝐵𝐿𝐺AM_{B}^{LG}italic_A italic_M start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_L italic_G end_POSTSUPERSCRIPT transits from a statistically more perpendicular alignment to a statistically more parallel alignment, then transit back to AMBLG0similar-to𝐴superscriptsubscript𝑀𝐵𝐿𝐺0AM_{B}^{LG}\sim 0italic_A italic_M start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_L italic_G end_POSTSUPERSCRIPT ∼ 0 as N𝑁Nitalic_N increases. As ϕBLGsuperscriptsubscriptitalic-ϕ𝐵𝐿𝐺\phi_{B}^{LG}italic_ϕ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_L italic_G end_POSTSUPERSCRIPT directly traces the correlation between magnetic fields and gravity, the observed AM-N trend for ϕBLGsuperscriptsubscriptitalic-ϕ𝐵𝐿𝐺\phi_{B}^{LG}italic_ϕ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_L italic_G end_POSTSUPERSCRIPT suggests that the magnetic field is resisting gravitational distortion at lower density, but is dragged by gravity as density increases within the cloud, and might be affected by early massive star formation activities near the central young stellar objects at even higher densities (Liu et al., 2023a).

Figure 14: Relative orientations (characterized by AM𝐴𝑀AMitalic_A italic_M) between magnetic field (θBsubscript𝜃B\theta_{\mathrm{B}}italic_θ start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT) and local gravity (θLGsubscript𝜃LG\theta_{\mathrm{LG}}italic_θ start_POSTSUBSCRIPT roman_LG end_POSTSUBSCRIPT) as a function of column density for Planck (left), JCMT (middle), and ALMA (right) observations.

3.6.3 Normalised mass-to-flux ratio

Based on ideal MHD equations, Koch et al. (2012b) suggested that the normalized mass-to-flux ratio can be estimated with

λKTH=ΣB1/2π1/2,subscript𝜆KTHdelimited-⟨⟩superscriptsubscriptΣ𝐵12superscript𝜋12\lambda_{\mathrm{KTH}}=\langle\Sigma_{B}^{-1/2}\rangle\pi^{-1/2},italic_λ start_POSTSUBSCRIPT roman_KTH end_POSTSUBSCRIPT = ⟨ roman_Σ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 / 2 end_POSTSUPERSCRIPT ⟩ italic_π start_POSTSUPERSCRIPT - 1 / 2 end_POSTSUPERSCRIPT , (21)

where ΣBsubscriptΣ𝐵\Sigma_{B}roman_Σ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT is the local ratio between the magnetic field force (FBsubscript𝐹𝐵F_{B}italic_F start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT) and the gravitational force (FGsubscript𝐹𝐺F_{G}italic_F start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT). ΣBsubscriptΣ𝐵\Sigma_{B}roman_Σ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT can be estimated with

ΣB=sinϕLGNGsin(90°ϕBNG)=FB|FG|subscriptΣ𝐵superscriptsubscriptitalic-ϕ𝐿𝐺𝑁𝐺90°superscriptsubscriptitalic-ϕ𝐵𝑁𝐺subscript𝐹𝐵subscript𝐹𝐺\Sigma_{B}=\frac{\sin\phi_{LG}^{NG}}{\sin(90\arcdeg-\phi_{B}^{NG})}=\frac{F_{B% }}{|F_{G}|}roman_Σ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT = divide start_ARG roman_sin italic_ϕ start_POSTSUBSCRIPT italic_L italic_G end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N italic_G end_POSTSUPERSCRIPT end_ARG start_ARG roman_sin ( 90 ° - italic_ϕ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N italic_G end_POSTSUPERSCRIPT ) end_ARG = divide start_ARG italic_F start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT end_ARG start_ARG | italic_F start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT | end_ARG (22)

if the hydrostatic gas pressure is negligible.

Figure 15 shows the λKTHsubscript𝜆KTH\lambda_{\mathrm{KTH}}italic_λ start_POSTSUBSCRIPT roman_KTH end_POSTSUBSCRIPT derived from the KTH method. For the Planck data, there is λKTH1similar-tosubscript𝜆KTH1\lambda_{\mathrm{KTH}}\sim 1italic_λ start_POSTSUBSCRIPT roman_KTH end_POSTSUBSCRIPT ∼ 1, which suggests a magnetically trans-critical state in the environment. For the JCMT data, the λKTHsubscript𝜆KTH\lambda_{\mathrm{KTH}}italic_λ start_POSTSUBSCRIPT roman_KTH end_POSTSUBSCRIPT increases with N𝑁Nitalic_N and transits from λKTH<1subscript𝜆KTH1\lambda_{\mathrm{KTH}}<1italic_λ start_POSTSUBSCRIPT roman_KTH end_POSTSUBSCRIPT < 1 to λKTH>1subscript𝜆KTH1\lambda_{\mathrm{KTH}}>1italic_λ start_POSTSUBSCRIPT roman_KTH end_POSTSUBSCRIPT > 1. The discrepancy between the Planck and JCMT data might be because the JCMT data filters out the large-scale emission and underestimates the gravitational force at lower column densities. For the ALMA data, overall there is λKTH1greater-than-or-equivalent-tosubscript𝜆KTH1\lambda_{\mathrm{KTH}}\gtrsim 1italic_λ start_POSTSUBSCRIPT roman_KTH end_POSTSUBSCRIPT ≳ 1. However, the magnetic field may be affected by star formation feedback from central young stellar objects, which could make the KTH method not applicable to the ALMA data. Note that the systematic uncertainty of the λKTHsubscript𝜆KTH\lambda_{\mathrm{KTH}}italic_λ start_POSTSUBSCRIPT roman_KTH end_POSTSUBSCRIPT value estimated with the KTH method is unclear due to the lack of numerical tests.

Figure 15: Normalised mass-to-flux ratio derived from the KTH method as a function of column density for Planck (left), JCMT (middle), and ALMA (right) observations.

4 Discussion

4.1 Equilibrium state

Different star formation theories (e.g., Bonnell et al., 1997; McKee & Tan, 2002) have distinct predictions over the energy balance between gravity, magnetic fields, and turbulence in different scales and evolutionary stages of massive star formation. Observationally determining the energy budget of massive star formation regions is key to distinguishing between those theories. Here we use the virial theorem to characterize the multi-scale equilibrium state of G28.34. Neglecting the surface term, tension term, and ordered velocity motion, the virial theorem is written as

12d2Idt2=EG+2Eth+2Eturb+EB,12superscript𝑑2𝐼𝑑superscript𝑡2subscript𝐸G2subscript𝐸th2subscript𝐸turbsubscript𝐸B\frac{1}{2}\frac{d^{2}I}{dt^{2}}=E_{\mathrm{G}}+2E_{\mathrm{th}}+2E_{\mathrm{% turb}}+E_{\mathrm{B}},divide start_ARG 1 end_ARG start_ARG 2 end_ARG divide start_ARG italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_I end_ARG start_ARG italic_d italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG = italic_E start_POSTSUBSCRIPT roman_G end_POSTSUBSCRIPT + 2 italic_E start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT + 2 italic_E start_POSTSUBSCRIPT roman_turb end_POSTSUBSCRIPT + italic_E start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT , (23)

where I𝐼Iitalic_I is the moment of inertia, EGsubscript𝐸GE_{\mathrm{G}}italic_E start_POSTSUBSCRIPT roman_G end_POSTSUBSCRIPT is the gravitational energy, Ethsubscript𝐸thE_{\mathrm{th}}italic_E start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT is the thermal energy, Eturbsubscript𝐸turbE_{\mathrm{turb}}italic_E start_POSTSUBSCRIPT roman_turb end_POSTSUBSCRIPT is the turbulent kinetic energy, and EBsubscript𝐸BE_{\mathrm{B}}italic_E start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT is the magnetic energy. 12d2Idt20greater-than-or-equivalent-to12superscript𝑑2𝐼𝑑superscript𝑡20\frac{1}{2}\frac{d^{2}I}{dt^{2}}\gtrsim 0divide start_ARG 1 end_ARG start_ARG 2 end_ARG divide start_ARG italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_I end_ARG start_ARG italic_d italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ≳ 0 indicates a supported state in Quasi-Equilibrium, while 12d2Idt2<012superscript𝑑2𝐼𝑑superscript𝑡20\frac{1}{2}\frac{d^{2}I}{dt^{2}}<0divide start_ARG 1 end_ARG start_ARG 2 end_ARG divide start_ARG italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_I end_ARG start_ARG italic_d italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG < 0 indicates a dynamical collapsing state. For a spherical structure with density profile nriproportional-to𝑛superscript𝑟𝑖n\propto r^{-i}italic_n ∝ italic_r start_POSTSUPERSCRIPT - italic_i end_POSTSUPERSCRIPT, the gravitational energy is given by EG=GM2/(kiR)subscript𝐸G𝐺superscript𝑀2subscript𝑘𝑖𝑅E_{\mathrm{G}}=-GM^{2}/(k_{i}R)italic_E start_POSTSUBSCRIPT roman_G end_POSTSUBSCRIPT = - italic_G italic_M start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ( italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_R ), where ki=(52i)/(3i)subscript𝑘𝑖52𝑖3𝑖k_{i}=(5-2i)/(3-i)italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = ( 5 - 2 italic_i ) / ( 3 - italic_i ), R𝑅Ritalic_R is the radius, and G𝐺Gitalic_G is the gravitational constant. The thermal energy is given by Eth=1.5nkBTVsubscript𝐸th1.5𝑛subscript𝑘𝐵𝑇𝑉E_{\mathrm{th}}=1.5nk_{B}TVitalic_E start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT = 1.5 italic_n italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T italic_V, where V=4πR3/3𝑉4𝜋superscript𝑅33V=4\pi R^{3}/3italic_V = 4 italic_π italic_R start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT / 3 is the Volume. The turbulent kinetic energy is given by Eturb=1.5Mσturb2subscript𝐸turb1.5𝑀superscriptsubscript𝜎𝑡𝑢𝑟𝑏2E_{\mathrm{turb}}=1.5M\sigma_{turb}^{2}italic_E start_POSTSUBSCRIPT roman_turb end_POSTSUBSCRIPT = 1.5 italic_M italic_σ start_POSTSUBSCRIPT italic_t italic_u italic_r italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, where σturbsubscript𝜎turb\sigma_{\mathrm{turb}}italic_σ start_POSTSUBSCRIPT roman_turb end_POSTSUBSCRIPT is the 1D turbulent velocity dispersion. The magnetic energy is given by EB=B2V/(2μ0)subscript𝐸Bsuperscript𝐵2𝑉2subscript𝜇0E_{\mathrm{B}}=B^{2}V/(2\mu_{0})italic_E start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT = italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_V / ( 2 italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ).

4.1.1 Gravity vs Thermal force

The relative importance between gravity and the thermal term in the virial theorem can be characterized with the ratio 2Eth/|EG|2subscript𝐸thsubscript𝐸G2E_{\mathrm{th}}/|E_{\mathrm{G}}|2 italic_E start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT / | italic_E start_POSTSUBSCRIPT roman_G end_POSTSUBSCRIPT |. Using the fitted mass profile Mr1.59proportional-to𝑀superscript𝑟1.59M\propto r^{1.59}italic_M ∝ italic_r start_POSTSUPERSCRIPT 1.59 end_POSTSUPERSCRIPT and density profile nr1.41proportional-to𝑛superscript𝑟1.41n\propto r^{-1.41}italic_n ∝ italic_r start_POSTSUPERSCRIPT - 1.41 end_POSTSUPERSCRIPT (see Section 3.2) and assuming a constant temperature (T=15𝑇15T=15italic_T = 15 K), we have derived this thermal-to-gravitational ratio as a function of radius.

Figure 16: (a). Thermal-to-gravitational ratio 2Eth/|EG|2subscript𝐸thsubscript𝐸G2E_{\mathrm{th}}/|E_{\mathrm{G}}|2 italic_E start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT / | italic_E start_POSTSUBSCRIPT roman_G end_POSTSUBSCRIPT | as a function of radius. The horizontal dashed line indicates a balance between the thermal term 2Eth2subscript𝐸th2E_{\mathrm{th}}2 italic_E start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT and gravitational term EGsubscript𝐸GE_{\mathrm{G}}italic_E start_POSTSUBSCRIPT roman_G end_POSTSUBSCRIPT in the virial equation. (b). Turbulent-to-gravitational ratio 2Eturb/|EG|=αturb2subscript𝐸turbsubscript𝐸Gsubscript𝛼turb2E_{\mathrm{turb}}/|E_{\mathrm{G}}|=\alpha_{\mathrm{turb}}2 italic_E start_POSTSUBSCRIPT roman_turb end_POSTSUBSCRIPT / | italic_E start_POSTSUBSCRIPT roman_G end_POSTSUBSCRIPT | = italic_α start_POSTSUBSCRIPT roman_turb end_POSTSUBSCRIPT (i.e., turbulent virial parameter) as a function of radius. Different colors represent different regions. Different symbols indicate different instruments. Close symbols indicate αturbsubscript𝛼turb\alpha_{\mathrm{turb}}italic_α start_POSTSUBSCRIPT roman_turb end_POSTSUBSCRIPT estimated with LOS velocity dispersion σvsubscript𝜎𝑣\sigma_{v}italic_σ start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT, while open symbols indicate αturbsubscript𝛼turb\alpha_{\mathrm{turb}}italic_α start_POSTSUBSCRIPT roman_turb end_POSTSUBSCRIPT estimated with velocity centroid dispersion σcsubscript𝜎𝑐\sigma_{c}italic_σ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. The horizontal dashed line indicates a balance between the turbulent term 2Eturb2subscript𝐸turb2E_{\mathrm{turb}}2 italic_E start_POSTSUBSCRIPT roman_turb end_POSTSUBSCRIPT and gravitational term EGsubscript𝐸GE_{\mathrm{G}}italic_E start_POSTSUBSCRIPT roman_G end_POSTSUBSCRIPT in the virial equation. (c). Magnetic-to-gravitational ratio EB/|EG|=1/λ2subscript𝐸Bsubscript𝐸G1superscript𝜆2E_{\mathrm{B}}/|E_{\mathrm{G}}|=1/\lambda^{2}italic_E start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT / | italic_E start_POSTSUBSCRIPT roman_G end_POSTSUBSCRIPT | = 1 / italic_λ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT as a function of radius. Different colors represent different regions. Different symbols indicate different instruments and modified DCF methods. The horizontal dashed line indicates a balance between magnetic energy and gravitational energy.

Figure 16(a) shows the derived thermal-to-gravitational ratio as a function of radius r𝑟ritalic_r. The thermal force is negligible compared to the gravitational force throughout the scales of our interest which is critical for star formation. This is reasonable because the thermal energy is usually less dominant compared to other forces in massive star formation (Tan et al., 2014). There is a trend that 2Eth/|EG|2subscript𝐸thsubscript𝐸G2E_{\mathrm{th}}/|E_{\mathrm{G}}|2 italic_E start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT / | italic_E start_POSTSUBSCRIPT roman_G end_POSTSUBSCRIPT | increases as r𝑟ritalic_r decreases. At a very small scale near the central Young Stellar Objects, the thermal force might surpass the gravitational force and establish an equilibrium state solely supported by thermal pressure. Because the power-law indexes for the mass profile and density profile may change at small scales, higher-resolution observations are required to study whether a small-scale equilibrium between thermal force and gravity could be achieved, and if so, at what scale this equilibrium might occur.

4.1.2 Gravity vs Turbulence

The relative importance between gravity and turbulence is usually characterized by the turbulent virial parameter αturb=2Eturb/|EG|=Mturb/Msubscript𝛼turb2subscript𝐸turbsubscript𝐸Gsubscript𝑀turb𝑀\alpha_{\mathrm{turb}}=2E_{\mathrm{turb}}/|E_{\mathrm{G}}|=M_{\mathrm{turb}}/Mitalic_α start_POSTSUBSCRIPT roman_turb end_POSTSUBSCRIPT = 2 italic_E start_POSTSUBSCRIPT roman_turb end_POSTSUBSCRIPT / | italic_E start_POSTSUBSCRIPT roman_G end_POSTSUBSCRIPT | = italic_M start_POSTSUBSCRIPT roman_turb end_POSTSUBSCRIPT / italic_M (Bertoldi & McKee, 1992). For a spherical structure with density profile nriproportional-to𝑛superscript𝑟𝑖n\propto r^{-i}italic_n ∝ italic_r start_POSTSUPERSCRIPT - italic_i end_POSTSUPERSCRIPT, the turbulent virial mass Mturbsubscript𝑀turbM_{\mathrm{turb}}italic_M start_POSTSUBSCRIPT roman_turb end_POSTSUBSCRIPT is given by

Mturb=3kiRσturb2G.subscript𝑀turb3subscript𝑘𝑖𝑅superscriptsubscript𝜎turb2𝐺M_{\mathrm{turb}}=\frac{3k_{i}R\sigma_{\mathrm{turb}}^{2}}{G}.italic_M start_POSTSUBSCRIPT roman_turb end_POSTSUBSCRIPT = divide start_ARG 3 italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_R italic_σ start_POSTSUBSCRIPT roman_turb end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_G end_ARG . (24)

αturb<1subscript𝛼turb1\alpha_{\mathrm{turb}}<1italic_α start_POSTSUBSCRIPT roman_turb end_POSTSUBSCRIPT < 1 suggests the turbulence cannot solely resist gravitational collapse (i.e., sub-virial), and vice versa.

We have calculated the turbulent virial parameter αturbsubscript𝛼turb\alpha_{\mathrm{turb}}italic_α start_POSTSUBSCRIPT roman_turb end_POSTSUBSCRIPT for the multi-scale structures of G28.34. For the density profile, we adopt i=1.41𝑖1.41i=1.41italic_i = 1.41 (see Section 3.2). For the clumps and cores, we adopt the mass estimated in Section 3.2. For the environmental gas, we adopt the mass extrapolated from the fitted mass-radius relation Mr1.59proportional-to𝑀superscript𝑟1.59M\propto r^{1.59}italic_M ∝ italic_r start_POSTSUPERSCRIPT 1.59 end_POSTSUPERSCRIPT (see Section 3.2). As discussed in Section 3.3, the LOS velocity dispersion σvsubscript𝜎𝑣\sigma_{v}italic_σ start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT tends to overestimate the actual 1D velocity dispersion at scale r𝑟ritalic_r because the LOS substructures are superposed and the LOS depth is larger than r𝑟ritalic_r, while the velocity centroid dispersion σcsubscript𝜎𝑐\sigma_{c}italic_σ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT tends to underestimate the actual 1D velocity dispersion at scale r𝑟ritalic_r due to the LOS averaging. As it is hard to derive the pure turbulent velocity dispersion corresponding to a specific scale (r𝑟ritalic_r) in molecular clouds, we adopt the LOS velocity dispersion σvsubscript𝜎𝑣\sigma_{v}italic_σ start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT as the upper limit of the 1D turbulent velocity dispersion and the velocity centroid dispersion σcsubscript𝜎𝑐\sigma_{c}italic_σ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT as the lower limit, with the assumption that the non-turbulent part of σcsubscript𝜎𝑐\sigma_{c}italic_σ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT is much weaker than the turbulent part. For the JCMT line data, we only adopt σvsubscript𝜎𝑣\sigma_{v}italic_σ start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT and σcsubscript𝜎𝑐\sigma_{c}italic_σ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT estimated from 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO (3-2) data as its critical density is closer to the clump densities.

Figure 16(b) shows the estimated turbulent virial parameter αturbsubscript𝛼turb\alpha_{\mathrm{turb}}italic_α start_POSTSUBSCRIPT roman_turb end_POSTSUBSCRIPT as a function of radius r𝑟ritalic_r. It is clear that the cores in MM4 and MM9 are sub-virial. For the clumps, the cloud, and the environmental gas, the αturbsubscript𝛼turb\alpha_{\mathrm{turb}}italic_α start_POSTSUBSCRIPT roman_turb end_POSTSUBSCRIPT estimated with σvsubscript𝜎𝑣\sigma_{v}italic_σ start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT is super-virial, while the αturbsubscript𝛼turb\alpha_{\mathrm{turb}}italic_α start_POSTSUBSCRIPT roman_turb end_POSTSUBSCRIPT estimated with σcsubscript𝜎𝑐\sigma_{c}italic_σ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT is sub-virial. Due to the uncertainties in the turbulent velocity dispersion, we are unable to determine whether these large-scale structures are turbulence-supported or not.

4.1.3 Gravity vs Magnetic fields

The relative significance of gravity and magnetic fields can be assessed using their energy ratio EB/|EG|subscript𝐸Bsubscript𝐸GE_{\mathrm{B}}/|E_{\mathrm{G}}|italic_E start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT / | italic_E start_POSTSUBSCRIPT roman_G end_POSTSUBSCRIPT |. Alternatively, this comparison is more usually expressed by the normalized mass-to-flux ratio λ𝜆\lambdaitalic_λ (Crutcher et al., 2004). For a spherical structure with density profile nriproportional-to𝑛superscript𝑟𝑖n\propto r^{-i}italic_n ∝ italic_r start_POSTSUPERSCRIPT - italic_i end_POSTSUPERSCRIPT, λ𝜆\lambdaitalic_λ is given by (Liu et al., 2022a)

λ=μH2mH1.5μ0πG/kiNH2B3d.𝜆subscript𝜇subscriptH2subscript𝑚H1.5subscript𝜇0𝜋𝐺subscript𝑘𝑖subscript𝑁subscriptH2subscript𝐵3𝑑\lambda=\mu_{\mathrm{H_{2}}}m_{\mathrm{H}}\sqrt{1.5\mu_{0}\pi G/k_{i}}\frac{N_% {\mathrm{H_{2}}}}{B_{3d}}.italic_λ = italic_μ start_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_m start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT square-root start_ARG 1.5 italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_π italic_G / italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG divide start_ARG italic_N start_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_ARG start_ARG italic_B start_POSTSUBSCRIPT 3 italic_d end_POSTSUBSCRIPT end_ARG . (25)

λ>1𝜆1\lambda>1italic_λ > 1 indicates magnetic fields cannot solely resist gravitational collapse (i.e., magnetically super-critical), and vice versa. For a spherical structure, the relation between λ𝜆\lambdaitalic_λ and the magnetic-to-gravitational energy ratio is EB/|EG|=1/λ2subscript𝐸Bsubscript𝐸G1superscript𝜆2E_{\mathrm{B}}/|E_{\mathrm{G}}|=1/\lambda^{2}italic_E start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT / | italic_E start_POSTSUBSCRIPT roman_G end_POSTSUBSCRIPT | = 1 / italic_λ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT.

With the magnetic field strengths estimated in Section 3.4, we have calculated the energy ratio EB/|EG|subscript𝐸Bsubscript𝐸GE_{\mathrm{B}}/|E_{\mathrm{G}}|italic_E start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT / | italic_E start_POSTSUBSCRIPT roman_G end_POSTSUBSCRIPT | as well as the normalized mass-to-flux ratio λ𝜆\lambdaitalic_λ for the environmental gas within 15-pc, the clumps, and MM4-core4. In the calculation, the POS total magnetic field B𝐵Bitalic_B is converted to the 3D total magnetic field B3dsubscript𝐵3𝑑B_{3d}italic_B start_POSTSUBSCRIPT 3 italic_d end_POSTSUBSCRIPT with the relation B3dB×1.25similar-tosubscript𝐵3𝑑𝐵1.25B_{3d}\sim B\times 1.25italic_B start_POSTSUBSCRIPT 3 italic_d end_POSTSUBSCRIPT ∼ italic_B × 1.25 (Liu et al., 2022b). Similarly, we adopt i=1.41𝑖1.41i=1.41italic_i = 1.41 for the density profile and adopt the extrapolated mass and density for the environmental gas. Table 2 summarizes the calculated λ𝜆\lambdaitalic_λ values.

Table 2: Summary of physical parameters
Region n𝑛nitalic_n N𝑁Nitalic_N r𝑟ritalic_r M𝑀Mitalic_M B𝐵Bitalic_B aaThe former value adopted the CY16 correction and the latter value adopted the Liu21 or Ost01 correction. λ𝜆\lambdaitalic_λ aaThe former value adopted the CY16 correction and the latter value adopted the Liu21 or Ost01 correction.
(cm33{}^{-3}start_FLOATSUPERSCRIPT - 3 end_FLOATSUPERSCRIPT) (cm22{}^{-2}start_FLOATSUPERSCRIPT - 2 end_FLOATSUPERSCRIPT) (pc) (Msubscript𝑀direct-productM_{\odot}italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT) (mG)
MM4-core4 1.1 ×\times× 1066{}^{6}start_FLOATSUPERSCRIPT 6 end_FLOATSUPERSCRIPT 1.5 ×\times× 102323{}^{23}start_FLOATSUPERSCRIPT 23 end_FLOATSUPERSCRIPT 0.053 43.0 0.29, 0.10 3.32, 9.32
MM4 6.3 ×\times× 1033{}^{3}start_FLOATSUPERSCRIPT 3 end_FLOATSUPERSCRIPT 2.6 ×\times× 102222{}^{22}start_FLOATSUPERSCRIPT 22 end_FLOATSUPERSCRIPT 1 1875.8 0.08, 0.12 2.09, 1.41
MM6 5.8 ×\times× 1033{}^{3}start_FLOATSUPERSCRIPT 3 end_FLOATSUPERSCRIPT 2.4 ×\times× 102222{}^{22}start_FLOATSUPERSCRIPT 22 end_FLOATSUPERSCRIPT 1 1711.6 0.07, 0.12 2.36, 1.28
MM9 5.7 ×\times× 1033{}^{3}start_FLOATSUPERSCRIPT 3 end_FLOATSUPERSCRIPT 2.4 ×\times× 102222{}^{22}start_FLOATSUPERSCRIPT 22 end_FLOATSUPERSCRIPT 1 1681.9 0.04, 0.08 3.85, 2.02
MM10 6.0 ×\times× 1033{}^{3}start_FLOATSUPERSCRIPT 3 end_FLOATSUPERSCRIPT 2.5 ×\times× 102222{}^{22}start_FLOATSUPERSCRIPT 22 end_FLOATSUPERSCRIPT 1 1789.2 0.06, 0.15 2.69, 1.04
Environment 1.4 ×\times× 1022{}^{2}start_FLOATSUPERSCRIPT 2 end_FLOATSUPERSCRIPT 8.6 ×\times× 102121{}^{21}start_FLOATSUPERSCRIPT 21 end_FLOATSUPERSCRIPT 15 133159.6 0.07, 0.27 0.76, 0.21
Figure 17: (a)-(c). Normalised mass-to-flux ratio derived from the DCF method as functions of radius, number density, and column density for Planck, JCMT, and ALMA observations toward G28.34. Different colors represent different regions. Different symbols indicate different instruments and modified methods. The horizontal dashed line indicates an energy balance between magnetic energy and gravitational energy (i.e., λ=1𝜆1\lambda=1italic_λ = 1). (d). Normalised mass-to-flux ratio derived from the compilation of DCF estimations of different star formation regions in the literature before June 2021 (reproduced with permission. See reviews in Liu et al., 2022a, b, and see references therein). Different colors represent different DCF variants and different symbols indicate different observational instruments (see Figure 3 of Liu et al., 2022b).

Figure 16(c) shows the energy ratio EB/|EG|subscript𝐸Bsubscript𝐸GE_{\mathrm{B}}/|E_{\mathrm{G}}|italic_E start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT / | italic_E start_POSTSUBSCRIPT roman_G end_POSTSUBSCRIPT | as a function of radius. As most previous studies have used λ𝜆\lambdaitalic_λ to compare magnetic fields with gravity, we also show the estimated λ𝜆\lambdaitalic_λ values as functions of r𝑟ritalic_r, n𝑛nitalic_n, and N𝑁Nitalic_N in Figures 17(a)-(c). We see that λ𝜆\lambdaitalic_λ increases with increasing density and decreases with increasing radius. This trend is consistent with that derived from the KTH method (see Figure 15). The environmental gas tends to be magnetically sub-critical, which means the magnetic field is strong enough to resist gravitational collapse in the large-scale diffuse gas. In the clumps/cores (as revealed by JCMT and ALMA), the state transits to super-critical, allowing gravitational collapse to happen. The transition of the magnetic critical state occurs at the cloud-to-clump scale. At the core scale, the magnetic support is notably weaker, even weaker than the thermal support (see Figures 16(a) and (c)). Note that although the magnetic field strengths estimated with the CY16 and Liu21/Ost01 corrections are different, the critical states derived with the two different corrections are consistent. Although it is hard to assess the uncertainty of the λ𝜆\lambdaitalic_λ values estimated from the modified DCF methods due to the many assumptions adopted, it is reasonable to think that the systematic uncertainties from those assumptions may shift λ𝜆\lambdaitalic_λ toward larger or smaller values, but does not significantly change the general trend of λ𝜆\lambdaitalic_λ. The trend of increasing λ𝜆\lambdaitalic_λ with density in G28.34 is in agreement with the λN𝜆𝑁\lambda-Nitalic_λ - italic_N trend from the previous DCF estimations in the literature (reviewed in Liu et al., 2022a, b), which is shown in Figure 17(d) for comparison. Note that the sample of the DCF compilation statistics in Liu et al. (2022a, b) are mostly from different regions, while our analysis presents the multi-scale magnetic critical state in an IRDC for the first time. The similarity between the multi-scale magnetic critical state in IRDC G28.34 and in other star formation regions may suggest that early massive star formation regions follow a similar evolution route as in other star formation regions of different evolutionary stages and masses. The general trend of λ𝜆\lambdaitalic_λ appears to be consistent with magnetic field-controlled star formation theories (Mouschovias et al., 2006), where magnetically sub-critical clouds gradually form super-critical substructures that collapse. The increase of λ𝜆\lambdaitalic_λ at higher densities may be due to the dissipation of magnetic flux (e.g., ambipolar diffusion or magnetic reconnection, Mouschovias & Ciolek, 1999; Lazarian & Vishniac, 1999), or mass accumulation along magnetic field lines.

4.1.4 Gravity vs Sum of competing forces

The total equilibrium state of star formation regions can be determined by comparing |EG|subscript𝐸G|E_{\mathrm{G}}|| italic_E start_POSTSUBSCRIPT roman_G end_POSTSUBSCRIPT | with the competing terms (2Eth+2Eturb+EB2subscript𝐸th2subscript𝐸turbsubscript𝐸B2E_{\mathrm{th}}+2E_{\mathrm{turb}}+E_{\mathrm{B}}2 italic_E start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT + 2 italic_E start_POSTSUBSCRIPT roman_turb end_POSTSUBSCRIPT + italic_E start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT) in the virial equation. Similar to the turbulent virial parameter, we could define a total virial parameter as the ratio of energies:

αtotal,E=2Eth+2Eturb+EB|EG|.subscript𝛼totalE2subscript𝐸th2subscript𝐸turbsubscript𝐸Bsubscript𝐸G\alpha_{\mathrm{total,E}}=\frac{2E_{\mathrm{th}}+2E_{\mathrm{turb}}+E_{\mathrm% {B}}}{|E_{\mathrm{G}}|}.italic_α start_POSTSUBSCRIPT roman_total , roman_E end_POSTSUBSCRIPT = divide start_ARG 2 italic_E start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT + 2 italic_E start_POSTSUBSCRIPT roman_turb end_POSTSUBSCRIPT + italic_E start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT end_ARG start_ARG | italic_E start_POSTSUBSCRIPT roman_G end_POSTSUBSCRIPT | end_ARG . (26)

Alternatively, the total virial parameter can be defined as the ratio between the total virial mass and mass:

αtotal,M=MtotalM.subscript𝛼totalMsubscript𝑀total𝑀\alpha_{\mathrm{total,M}}=\frac{M_{\mathrm{total}}}{M}.italic_α start_POSTSUBSCRIPT roman_total , roman_M end_POSTSUBSCRIPT = divide start_ARG italic_M start_POSTSUBSCRIPT roman_total end_POSTSUBSCRIPT end_ARG start_ARG italic_M end_ARG . (27)

Note that the values of αtotal,Esubscript𝛼totalE\alpha_{\mathrm{total,E}}italic_α start_POSTSUBSCRIPT roman_total , roman_E end_POSTSUBSCRIPT and αtotal,Msubscript𝛼totalM\alpha_{\mathrm{total,M}}italic_α start_POSTSUBSCRIPT roman_total , roman_M end_POSTSUBSCRIPT are not equal. The total virial mass is given by (Liu et al., 2020)

Mtotal=MB2+(Mturb+Mth2)2+Mturb+Mth2,subscript𝑀totalsubscriptsuperscript𝑀2Bsuperscriptsubscript𝑀turbsubscript𝑀th22subscript𝑀turbsubscript𝑀th2M_{\mathrm{total}}=\sqrt{M^{2}_{\mathrm{B}}+(\frac{M_{\mathrm{turb}}+M_{% \mathrm{th}}}{2})^{2}}+\frac{M_{\mathrm{turb}}+M_{\mathrm{th}}}{2},italic_M start_POSTSUBSCRIPT roman_total end_POSTSUBSCRIPT = square-root start_ARG italic_M start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT + ( divide start_ARG italic_M start_POSTSUBSCRIPT roman_turb end_POSTSUBSCRIPT + italic_M start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + divide start_ARG italic_M start_POSTSUBSCRIPT roman_turb end_POSTSUBSCRIPT + italic_M start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG , (28)

where the thermal virial mass is estimated with

Mth=3kiRGkBTμH2mH,subscript𝑀th3subscript𝑘𝑖𝑅𝐺subscript𝑘𝐵𝑇subscript𝜇subscriptH2subscript𝑚HM_{\mathrm{th}}=\frac{3k_{i}R}{G}\frac{k_{B}T}{\mu_{\mathrm{H_{2}}}m_{\mathrm{% H}}},italic_M start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT = divide start_ARG 3 italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_R end_ARG start_ARG italic_G end_ARG divide start_ARG italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T end_ARG start_ARG italic_μ start_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_m start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT end_ARG , (29)

and the magnetic virial mass444Some previous studies wrote the magnetic virial mass of a uniform spherical structure as MB=(5VA2R)/(6G)subscriptsuperscript𝑀B5superscriptsubscript𝑉A2𝑅6𝐺M^{\prime}_{\mathrm{B}}=(5V_{\mathrm{A}}^{2}R)/(6G)italic_M start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT = ( 5 italic_V start_POSTSUBSCRIPT roman_A end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_R ) / ( 6 italic_G ), where VAsubscript𝑉AV_{\mathrm{A}}italic_V start_POSTSUBSCRIPT roman_A end_POSTSUBSCRIPT is the 3D Alfvénic velocity. As demonstrated in the Appendix of Liu et al. (2020), MBsubscriptsuperscript𝑀BM^{\prime}_{\mathrm{B}}italic_M start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT underestimates MBsubscript𝑀BM_{\mathrm{B}}italic_M start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT when M>MB𝑀subscript𝑀BM>M_{\mathrm{B}}italic_M > italic_M start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT, and vice versa. Thus, MBsubscriptsuperscript𝑀BM^{\prime}_{\mathrm{B}}italic_M start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT is wrong and should not be used in the virial calculation. is estimated with

MB=πR2B1.5μ0πG/ki.subscript𝑀B𝜋superscript𝑅2𝐵1.5subscript𝜇0𝜋𝐺subscript𝑘𝑖M_{\mathrm{B}}=\frac{\pi R^{2}B}{\sqrt{1.5\mu_{0}\pi G/k_{i}}}.italic_M start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT = divide start_ARG italic_π italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_B end_ARG start_ARG square-root start_ARG 1.5 italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_π italic_G / italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG end_ARG . (30)

The environmental gas should have αtotal>1subscript𝛼total1\alpha_{\mathrm{total}}>1italic_α start_POSTSUBSCRIPT roman_total end_POSTSUBSCRIPT > 1 since the magnetic field itself is stronger than gravity. i.e., gravity is unimportant in the large-scale diffuse gas. At the intermediate scale, although it is clear that magnetic fields are dominated by gravity in the clumps (i.e., λ>1𝜆1\lambda>1italic_λ > 1), it is hard to assess their total equilibrium states (considering both the turbulent and magnetic supports). This is because there are some uncertainties in the derivation of the pure turbulent velocity dispersion at specific scales for the single-dish data (see Section 3.3). Approximately adopting an average value between the LOS velocity dispersion and the velocity centroid dispersion as the turbulent velocity dispersion, we obtain that some clumps have αtotal<1subscript𝛼total1\alpha_{\mathrm{total}}<1italic_α start_POSTSUBSCRIPT roman_total end_POSTSUBSCRIPT < 1 while some clumps have αtotal>1subscript𝛼total1\alpha_{\mathrm{total}}>1italic_α start_POSTSUBSCRIPT roman_total end_POSTSUBSCRIPT > 1, with the average state being near quasi-equilibrium (i.e., αtotal¯¯subscript𝛼total\overline{\alpha_{\mathrm{total}}}over¯ start_ARG italic_α start_POSTSUBSCRIPT roman_total end_POSTSUBSCRIPT end_ARG not far from 1). Alternatively, assuming that the magnetic support is comparable to the turbulent support at the clump scale, we obtain similar results that the average state of clumps is near quasi-equilibrium. At the core scale, even when considering the upper limit of supporting forces, the total virial parameter for core MM4-core4 is only αtotal,E0.5similar-tosubscript𝛼totalE0.5\alpha_{\mathrm{total,E}}\sim 0.5italic_α start_POSTSUBSCRIPT roman_total , roman_E end_POSTSUBSCRIPT ∼ 0.5. Thus, it is safe to say that gravity dominates over the combination of competing forces for MM4-core4.

In summary, we suggest that the G28.34 cloud is located in a globally-supported environment and its clumps are likely in an approximate quasi-equilibrium state, but the cores therein are undergoing dynamic collapse.

4.1.5 Implications on massive star formation

It has been long debated whether molecular clouds and their substructures are in equilibrium or not. The Planck dust polarization survey (Planck Collaboration et al., 2016) has found that the Gould Belt Clouds, including one massive star formation region (Orion), are in magnetically sub-critical (i.e., λ<1𝜆1\lambda<1italic_λ < 1) states with DCF estimations. Later, Liu et al. (2023a) analyzed the Planck dust polarization data in another massive star formation region NGC 6334 and found that this region is also in a magnetically sub-critical state at large scale. Combined with our finding of magnetically sub-critical state in the environment of IRDC G28.34, we suggest that massive star-forming clouds may be globally supported by magnetic fields both in the early and late evolutionary stages. The quasi-equilibrium state of clouds is essential to explain the low star formation rate observed and to allow ambipolar diffusion to happen (McKee & Ostriker, 2007), but contradicts the idea that clouds are short-lived and are undergoing global dynamical collapse (i.e., the global hierarchical collapse model, Vázquez-Semadeni et al., 2019). On the other hand, the sub-critical state does not necessarily mean the region will expand or disperse. Sub-critical clouds could still create overdense regions via other mechanisms such as large-scale turbulent inertial flows (i.e., the inertial-flow model, Padoan et al., 2020) or local infall through magnetic channels (Koch et al., 2018, 2022), instead of via symmetric gravitational collapse.

Within the G28.34 cloud, while the clumps may not be far from equilibrium, the cores are dominated by gravity. This is consistent with the findings of previous studies that gravity is more important in higher-density regions, and that cores in both early and evolved massive star formation regions tend to be averagely dominated by gravity even considering both the magnetic and turbulent supports (see reviews in Liu et al., 2022a, b). The dynamic state of cores is inconsistent with the turbulent core accretion model (McKee & Tan, 2002), which predicts αtotal1similar-tosubscript𝛼total1\alpha_{\mathrm{total}}\sim 1italic_α start_POSTSUBSCRIPT roman_total end_POSTSUBSCRIPT ∼ 1 across different scales. The near-equilibrium state of clumps tends to be inconsistent with the competitive accretion model (Bonnell et al., 1997), which requires αtotal<1subscript𝛼total1\alpha_{\mathrm{total}}<1italic_α start_POSTSUBSCRIPT roman_total end_POSTSUBSCRIPT < 1 for cloud substructures (Krumholz et al., 2005). Thus, both the two major massive star formation models may need some modifications to be consistent with the observational results.

4.2 Magnetic fields or turbulence?

Different star formation theories have distinct explanations for the controlling factor of the formation and evolution of molecular clouds and cloud substructures: some emphasize the role of magnetic fields (Mouschovias et al., 2006), while others highlight the role of turbulence (Mac Low & Klessen, 2004). Qualitative and quantitative comparisons between magnetic fields and turbulence are needed to determine their relative importance in star formation.

There is no way to observationally compare the turbulent kinetic energy and the turbulent magnetic energy yet (Liu et al., 2022b). People usually assume an equipartition between turbulent kinetic and magnetic energies, which implicitly assumes that the total magnetic energy is larger than the turbulent kinetic energy. On the other hand, the relative importance between the ordered magnetic field and the turbulence can be directly derived from the dust polarization maps, without the need for molecular line observations. The relation between the 3D Alfvénic Mach number and the POS angular dispersion is A(ft/fu/Qc/fo)σθsimilar-tosubscript𝐴subscript𝑓𝑡subscript𝑓𝑢subscript𝑄𝑐subscript𝑓𝑜subscript𝜎𝜃\mathcal{M}_{A}\sim(f_{t}/f_{u}/Q_{c}/f_{o})\sigma_{\theta}caligraphic_M start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ∼ ( italic_f start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT / italic_f start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT / italic_Q start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_f start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT ) italic_σ start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT (Liu et al., 2022b), where fusubscript𝑓𝑢f_{u}italic_f start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT and ftsubscript𝑓𝑡f_{t}italic_f start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT are factors for the 3D-to-POS conversion of B0subscript𝐵0B_{0}italic_B start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and Btsubscript𝐵tB_{\mathrm{t}}italic_B start_POSTSUBSCRIPT roman_t end_POSTSUBSCRIPT, fosubscript𝑓𝑜f_{o}italic_f start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT is a correction factor for the ordered field contribution, and Qcsubscript𝑄𝑐Q_{c}italic_Q start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT is a correction factor to account for other effects (e.g., the LOS signal integration, the difference between the orientation and direction, et al.). However, the actual value of ftsubscript𝑓𝑡f_{t}italic_f start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT, fusubscript𝑓𝑢f_{u}italic_f start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT, and fosubscript𝑓𝑜f_{o}italic_f start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT is usually unclear in individual regions. Statistical values exist for the correction factors (e.g., Crutcher et al., 2004; Liu et al., 2021), but those statistical values (especially for fusubscript𝑓𝑢f_{u}italic_f start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT and fosubscript𝑓𝑜f_{o}italic_f start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT) may only be appropriate for statistical studies (e.g., Liu et al., 2022a; Pattle et al., 2023). Thus, we refrain from deriving Asubscript𝐴\mathcal{M}_{A}caligraphic_M start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT from polarization angular dispersions in G28.34.

The relative orientation analysis (Section 3.6) offers an alternative way to qualitatively compare magnetic fields and turbulence. With an approach equivalent to the HRO analysis, we find that the magnetic field and column density gradient transits from statistically more perpendicular to more parallel as column density increases for the Planck and JCMT observations (Section 3.6.1), which is a sign of trans-to-sub-Alfvénic turbulence at large scales. This transition of alignment is consistent with previous studies in low-mass star formation regions and in evolved massive star formation regions (Planck Collaboration et al., 2016; Liu et al., 2023a).

The statistical tool VGT offers another way to compare magnetic fields and turbulence at large scales. Implementing the VGT, we find A=0.74subscript𝐴0.74\mathcal{M}_{A}=0.74caligraphic_M start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT = 0.74 within r=15𝑟15r=15italic_r = 15 pc, which implies that G28.34 is located in a slightly sub-Alfvénic environment. The sub-Alfvénic state is consistent with the results from previous VGT studies in nearby low-mass star formation regions (Hu et al., 2019).

In summary, we conclude that both low-mass and high-mass star formation happens in a trans-to-sub-Alfvénic environment, which means magnetic fields play a more important role than turbulence in controlling star formation at large scales. Since gravity is not important in large-scale diffuse gas, cloud formation and evolution at the cloud scale should be mainly controlled by the property of trans-to-sub-Alfvénic MHD turbulence. Within the cloud, the situation is more complicated at small scales. More reliable analysis methods are required to compare magnetic fields and turbulence in high-density and small-scale regions with significant gravity.

5 Summary

With JCMT and ALMA dust polarization observations as well as Planck dust polarization data, we present a study of the multi-scale magnetic fields in IRDC G28.34. We also have studied the multi-scale velocity fields in G28.34 with molecular line data from FCRAO-14m, JCMT, and ALMA. The findings are:

  1. 1.

    Within our JCMT detection region, 5 clumps (MM1, MM2, MM11, MM16, and MM17) have average magnetic fields aligned within 30°°\arcdeg° of the cloud-scale magnetic field, while 5 clumps (MM4, MM6, MM9, MM10, and MM14) have averaged magnetic fields misaligned at 60°°\arcdeg°-90°°\arcdeg° with respect to the cloud-scale magnetic field. The bimodal distribution suggests that the clump-scale magnetic field is organized with respect to the cloud-scale magnetic field. In MM4, the core-to-condensation scale magnetic field is preferentially aligned with the clump-scale magnetic field and the clump-scale magnetic field is perpendicular to the chain of fragments. This may suggest that the magnetic field plays a crucial role in the clump collapse and fragmentation process.

  2. 2.

    With a simple power-law fit for the mass-radius and density-radius relations of G28.34, we obtain Mr1.59proportional-to𝑀superscript𝑟1.59M\propto r^{1.59}italic_M ∝ italic_r start_POSTSUPERSCRIPT 1.59 end_POSTSUPERSCRIPT and nr1.41proportional-to𝑛superscript𝑟1.41n\propto r^{-1.41}italic_n ∝ italic_r start_POSTSUPERSCRIPT - 1.41 end_POSTSUPERSCRIPT between similar-to\sim0.07 pc and similar-to\sim7 pc.

  3. 3.

    We have studied the multi-scale relative orientations between magnetic fields, column density gradients, and local gravity in G28.34. As column densities increase, the magnetic field and column density gradient transit from preferentially more perpendicular to more parallel, then transit back to a random alignment. The alignment between the magnetic field and local gravity shows a similar varying trend with column density. The results of the relative orientation analysis suggest that G28.34 is located in a trans-to-sub-Alfvénic environment, the magnetic field is resisting gravitational collapse in the large-scale diffuse gas, the magnetic field is distorted by gravity within the cloud, and the magnetic field is affected by star formation activities in high-density regions.

  4. 4.

    We have measured the magnetic field strengths in the environmental gas, infrared dark clumps, and core MM4-core4 of G28.34 with modified DCF analysis. In this early massive star formation region, the magnetic field strengths do not significantly increase as density increases. With the estimated magnetic field strength, we find that the normalized mass-to-flux ratio λ𝜆\lambdaitalic_λ increases with density but decreases with radius, and transits from magnetically subcritical (λ<1𝜆1\lambda<1italic_λ < 1) in the environmental gas to supercritical (λ<1𝜆1\lambda<1italic_λ < 1) at clump/core scales. This is in agreement with the prediction of magnetic field-controlled star formation theories. With an alternative analysis using the KTH method, we find a similar increasing trend of λ𝜆\lambdaitalic_λ with density, except for the Planck data where λ1similar-to𝜆1\lambda\sim 1italic_λ ∼ 1.

  5. 5.

    Combining the thermal, turbulent, and magnetic supports, we find that the environmental gas is super-virial (αtotal>1subscript𝛼total1\alpha_{\mathrm{total}}>1italic_α start_POSTSUBSCRIPT roman_total end_POSTSUBSCRIPT > 1, supported) and MM4-core4 is sub-virial (αtotal<1subscript𝛼total1\alpha_{\mathrm{total}}<1italic_α start_POSTSUBSCRIPT roman_total end_POSTSUBSCRIPT < 1, gravity-dominant). The infrared dark clumps may be averagely in a near-equilibrium state (αtotal1similar-tosubscript𝛼total1\alpha_{\mathrm{total}}\sim 1italic_α start_POSTSUBSCRIPT roman_total end_POSTSUBSCRIPT ∼ 1). The transition from αtotal>1subscript𝛼total1\alpha_{\mathrm{total}}>1italic_α start_POSTSUBSCRIPT roman_total end_POSTSUBSCRIPT > 1 to αtotal<1subscript𝛼total1\alpha_{\mathrm{total}}<1italic_α start_POSTSUBSCRIPT roman_total end_POSTSUBSCRIPT < 1 is inconsistent with either the competitive accretion model or the turbulent core accretion model, suggesting that the two major massive star formation models may need some modifications.

  6. 6.

    With a VGT analysis, we find A=0.74subscript𝐴0.74\mathcal{M}_{A}=0.74caligraphic_M start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT = 0.74 within r=15𝑟15r=15italic_r = 15 pc centered at G28.34. The slightly sub-Alfvénic state is consistent with the relative orientation analysis and implies that magnetic fields regulate cloud formation and evolution at large scales. More reliable analysis methods are required to compare magnetic fields and turbulence in high-density regions where gravity is dominant.

We thank the anonymous referee for the constructive and in-depth comments that have improved the clarity of this work, especially on the discussions of the large-scale LOS contamination, the turbulence anisotropy, and the small-scale thermal effects. J.L. thanks Prof. Hua-Bai Li for helpful comments on the corresponding scales of linewidth and Ms. Yuchen Xing for helpful discussions on the mass-size relation. J.L. also thanks Prof. Patricio Sanhueza, Dr. Paulo Cortes, and the active JCMT BISTRO team members for helpful general discussions on magnetic field properties.

J.L. is partially supported by a Grant-in-Aid for Scientific Research (KAKENHI Number JP23H01221) of JSPS and was supported by the EAO Fellowship Program under the umbrella of the East Asia Core Observatories Association. K.Q. is supported by National Key R&D Program of China grant No. 2022YFA1603100. K.Q. acknowledges the support from National Natural Science Foundation of China (NSFC) through grant Nos. U1731237, 11590781, and 11629302. H.B.L. is supported by the National Science and Technology Council (NSTC) of Taiwan (Grant Nos. 111-2112-M-110-022-MY3). Z.Y.L. is supported in part by NSF AST-2307199 and NASA 80NSSC20K0533. J.M.G. acknowledges the support from the program Unidad de Excelencia María de Maeztu CEX2020-001058-M and the grant PID2020-117710GB-I00 (MCI-AEI-FEDER, UE). TGSP gratefully acknowledges support by the National Science Foundation under grant No. AST-2009842 and AST-2108989.

This paper makes use of the following ALMA data: ADS/JAO.ALMA#2016.1.00248.S and ADS/JAO.ALMA#2017.1.00793.S. ALMA is a partnership of the ESO (representing its member states), NSF (USA) and NINS (Japan), together with NRC (Canada), MOST and ASIAA (Taiwan), and KASI (Republic of Korea), in cooperation with the Republic of Chile. The Joint ALMA Observatory is operated by ESO, AUI/NRAO, and NAOJ. The National Radio Astronomy Observatory is a facility of the National Science Foundation operated under cooperative agreement by Associated Universities, Inc. The JCMT is operated by the EAO on behalf of NAOJ; ASIAA; KASI; CAMS as well as the National Key R&D Program of China (No. 2017YFA0402700). Additional funding support is provided by the STFC and participating universities in the UK and Canada. Additional funds for the construction of SCUBA-2 were provided by the Canada Foundation for Innovation. This work is based on observations obtained with Planck (http://www.esa.int/Planck), an ESA science mission with instruments and contributions directly funded by ESA Member States, NASA, and Canada.

Appendix A Polarization percentage

Figure 18 shows the polarization percentage map of our POL-2 data. Figure 19 shows the PI𝑃𝐼P-Iitalic_P - italic_I relation. The polarization percentage clearly decreases with increasing intensity. A small portion of data points have P>15%𝑃percent15P>15\%italic_P > 15 %, but previous theoretical models suggested that the polarization percentage of submm observations should not exceed 15% (Draine & Fraisse, 2009). We conservatively excluded data points with P>15%𝑃percent15P>15\%italic_P > 15 % in our analysis. Note that the large-scale extended emissions are filtered out during the JCMT data reduction processes. The I𝐼Iitalic_I emission may be more extended than the polarized emission in G28.34. Thus, the JCMT data reduction may have removed more I𝐼Iitalic_I emissions than Q𝑄Qitalic_Q and U𝑈Uitalic_U, which could lead to a systematical overestimation of P𝑃Pitalic_P. This presents a plausible explanation for the high P𝑃Pitalic_P values observed in weak-emission regions.

Figure 18: Polarization percentage (colorscales) of our POL-2 data (S/N(PI𝑃𝐼PIitalic_P italic_I)>>>2). Black solid contours indicate the JCMT Stokes-I intensities. Contour starts at 50 mJy beam11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT and continues at 150 mJy beam11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT.
Figure 19: Relation between polarization percentage and total intensity for our POL-2 data (S/N(I𝐼Iitalic_I)>>>3) within the central r=3\arcmin region. Grey and black colors correspond to 2<<<S/N(PI𝑃𝐼PIitalic_P italic_I)<<<3 and S/N(PI𝑃𝐼PIitalic_P italic_I)>>>3, respectively. The horizontal dashed line indicates P=15𝑃15P=15italic_P = 15%. The solid line indicates the result from a simple power-law fit for the PI𝑃𝐼P-Iitalic_P - italic_I relation for data points with S/N(PI𝑃𝐼PIitalic_P italic_I)>>>3.

Appendix B Molecular line

Figures 20 and 21 show the integrated intensity maps of the FCRAO-14m 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO (1-0), JCMT 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO (3-2) and HCO+{}^{+}start_FLOATSUPERSCRIPT + end_FLOATSUPERSCRIPT (4-3), and ALMA N22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPTD+{}^{+}start_FLOATSUPERSCRIPT + end_FLOATSUPERSCRIPT (3-2) data. The integrated velocity ranges are indicated in each panel.

Figure 20: (a) Integrated intensity of FCRAO-14m 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO (1-0) in the surrounding of G28.34. Black contours correspond to the Planck τ353subscript𝜏353\tau_{353}italic_τ start_POSTSUBSCRIPT 353 end_POSTSUBSCRIPT map, starting from 0.0005 and continuing with an interval of 0.0005. A dashed line indicates the galactic plane (b=0°𝑏0°b=0\arcdegitalic_b = 0 °). The dashed rectangle indicates the JCMT map area in (b) and (c). (b)-(c) Integrated intensity of JCMT 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO (3-2) and HCO+{}^{+}start_FLOATSUPERSCRIPT + end_FLOATSUPERSCRIPT (4-3) toward the IRDC G28.34. Black contours correspond to the JCMT 0.85 mm dust continuum map. Contour starts at 50 mJy beam11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT and continues at 150 mJy beam11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT.

Figure 21: Integrated intensity of ALMA N22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPTD+{}^{+}start_FLOATSUPERSCRIPT + end_FLOATSUPERSCRIPT (3-2) toward MM4 and MM9. Contours correspond to the ALMA 1.3 mm dust continuum map. Contour levels are (±plus-or-minus\pm±3, 6, 10, 20, 30, 40, 50, 70, 90, 110, 150, 180, 210, 250, 290, 340, 390, 450) ×σIabsentsubscript𝜎𝐼\times\sigma_{I}× italic_σ start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT.

Figures 22, 23, and 24 show examples of the average spectra for different line data. Many lines show signs of multiple velocity components, which provides evidence for the superposition of substructures along the LOS. We do not try to identify or separate the multiple velocity components because the combination of those components could appear as a single Gaussian line profile at a coarser spatial resolution, and there are always multiple velocity components seen with higher resolution observations inside a “single” velocity component. We perform a single-peak Gaussian fit for each line and obtain the best-fitting result. The fitted velocity dispersion should be the upper limit of the pure turbulent velocity dispersion.

Figure 22: Example of the average FCRAO-14m 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO (1-0) spectra within the central 15-pc (radius) area. The dashed single-Gaussian profile indicates the best-fitting result. The vertical dashed lines indicate the velocity ranges used for the analyses.
Figure 23: Example of the average JCMT 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO (3-2) (black) and HCO+{}^{+}start_FLOATSUPERSCRIPT + end_FLOATSUPERSCRIPT (4-3) (red) spectra of each clump within the central 1-pc (radius) area. The dashed Gaussian profiles indicate the best-fitting results for 1313{}^{13}start_FLOATSUPERSCRIPT 13 end_FLOATSUPERSCRIPTCO (3-2). The vertical dashed lines indicate the velocity ranges used for the analyses.
Figure 24: Average ALMA N22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPTD+{}^{+}start_FLOATSUPERSCRIPT + end_FLOATSUPERSCRIPT (3-2) spectra of MM4 and MM9 within the whole emission area in the ALMA field. The dashed Gaussian profiles indicate the best-fitting results. The vertical dashed lines indicate the velocity ranges used for the analyses. Our computing ability for ALMA data reduction is limited, so we only imaged the line within velocity ranges from 75 to 83 km s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT.

References

  • Añez-López et al. (2020) Añez-López, N., Busquet, G., Koch, P. M., et al. 2020, A&A, 644, A52, doi: 10.1051/0004-6361/202039152
  • Andersson et al. (2015) Andersson, B. G., Lazarian, A., & Vaillancourt, J. E. 2015, ARA&A, 53, 501, doi: 10.1146/annurev-astro-082214-122414
  • Astropy Collaboration et al. (2013) Astropy Collaboration, Robitaille, T. P., Tollerud, E. J., et al. 2013, A&A, 558, A33, doi: 10.1051/0004-6361/201322068
  • Astropy Collaboration et al. (2018) Astropy Collaboration, Price-Whelan, A. M., Sipőcz, B. M., et al. 2018, AJ, 156, 123, doi: 10.3847/1538-3881/aabc4f
  • Beattie et al. (2022) Beattie, J. R., Krumholz, M. R., Skalidis, R., et al. 2022, MNRAS, 515, 5267, doi: 10.1093/mnras/stac2099
  • Bergin & Tafalla (2007) Bergin, E. A., & Tafalla, M. 2007, ARA&A, 45, 339, doi: 10.1146/annurev.astro.45.071206.100404
  • Bertoldi & McKee (1992) Bertoldi, F., & McKee, C. F. 1992, ApJ, 395, 140, doi: 10.1086/171638
  • Beuther et al. (2007) Beuther, H., Leurini, S., Schilke, P., et al. 2007, A&A, 466, 1065, doi: 10.1051/0004-6361:20066742
  • Beuther et al. (2018) Beuther, H., Soler, J. D., Vlemmings, W., et al. 2018, A&A, 614, A64, doi: 10.1051/0004-6361/201732378
  • Beuther et al. (2020) Beuther, H., Wang, Y., Soler, J., et al. 2020, A&A, 638, A44, doi: 10.1051/0004-6361/202037950
  • Bonnell et al. (1997) Bonnell, I. A., Bate, M. R., Clarke, C. J., & Pringle, J. E. 1997, MNRAS, 285, 201, doi: 10.1093/mnras/285.1.201
  • Buckle et al. (2009) Buckle, J. V., Hills, R. E., Smith, H., et al. 2009, MNRAS, 399, 1026, doi: 10.1111/j.1365-2966.2009.15347.x
  • Carey et al. (1998) Carey, S. J., Clark, F. O., Egan, M. P., et al. 1998, ApJ, 508, 721, doi: 10.1086/306438
  • Chandrasekhar & Fermi (1953) Chandrasekhar, S., & Fermi, E. 1953, ApJ, 118, 113, doi: 10.1086/145731
  • Chen et al. (2007) Chen, H.-R., Su, Y.-N., Liu, S.-Y., et al. 2007, ApJ, 654, L87, doi: 10.1086/510715
  • Cho & Yoo (2016) Cho, J., & Yoo, H. 2016, ApJ, 821, 21, doi: 10.3847/0004-637X/821/1/21
  • Churchwell et al. (2009) Churchwell, E., Babler, B. L., Meade, M. R., et al. 2009, PASP, 121, 213, doi: 10.1086/597811
  • Corradi et al. (1998) Corradi, R. L. M., Aznar, R., & Mampaso, A. 1998, MNRAS, 297, 617, doi: 10.1046/j.1365-8711.1998.01532.x
  • Cortes et al. (2019) Cortes, P. C., Hull, C. L. H., Girart, J. M., et al. 2019, ApJ, 884, 48, doi: 10.3847/1538-4357/ab378d
  • Crutcher et al. (2004) Crutcher, R. M., Nutter, D. J., Ward-Thompson, D., & Kirk, J. M. 2004, ApJ, 600, 279, doi: 10.1086/379705
  • Currie et al. (2014) Currie, M. J., Berry, D. S., Jenness, T., et al. 2014, in Astronomical Society of the Pacific Conference Series, Vol. 485, Astronomical Data Analysis Software and Systems XXIII, ed. N. Manset & P. Forshay, 391
  • Davis (1951) Davis, L. 1951, Physical Review, 81, 890, doi: 10.1103/PhysRev.81.890.2
  • Dickman & Kleiner (1985) Dickman, R. L., & Kleiner, S. C. 1985, ApJ, 295, 479, doi: 10.1086/163391
  • Draine & Fraisse (2009) Draine, B. T., & Fraisse, A. A. 2009, ApJ, 696, 1, doi: 10.1088/0004-637X/696/1/1
  • Federrath (2016) Federrath, C. 2016, Journal of Plasma Physics, 82, 535820601, doi: 10.1017/S0022377816001069
  • Friberg et al. (2016) Friberg, P., Bastien, P., Berry, D., et al. 2016, in Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, Vol. 9914, Millimeter, Submillimeter, and Far-Infrared Detectors and Instrumentation for Astronomy VIII, ed. W. S. Holland & J. Zmuidzinas, 991403, doi: 10.1117/12.2231943
  • Goldreich & Sridhar (1995) Goldreich, P., & Sridhar, S. 1995, ApJ, 438, 763, doi: 10.1086/175121
  • Goldsmith & Langer (1999) Goldsmith, P. F., & Langer, W. D. 1999, ApJ, 517, 209, doi: 10.1086/307195
  • Gómez et al. (2018) Gómez, G. C., Vázquez-Semadeni, E., & Zamora-Avilés, M. 2018, MNRAS, 480, 2939, doi: 10.1093/mnras/sty2018
  • Goodman et al. (1998) Goodman, A. A., Barranco, J. A., Wilner, D. J., & Heyer, M. H. 1998, ApJ, 504, 223, doi: 10.1086/306045
  • Hildebrand (1983) Hildebrand, R. H. 1983, QJRAS, 24, 267
  • Hildebrand et al. (1984) Hildebrand, R. H., Dragovan, M., & Novak, G. 1984, ApJ, 284, L51, doi: 10.1086/184351
  • Hildebrand et al. (2009) Hildebrand, R. H., Kirby, L., Dotson, J. L., Houde, M., & Vaillancourt, J. E. 2009, ApJ, 696, 567, doi: 10.1088/0004-637X/696/1/567
  • Holland et al. (2013) Holland, W. S., Bintley, D., Chapin, E. L., et al. 2013, MNRAS, 430, 2513, doi: 10.1093/mnras/sts612
  • Houde et al. (2016) Houde, M., Hull, C. L. H., Plambeck, R. L., Vaillancourt, J. E., & Hildebrand, R. H. 2016, ApJ, 820, 38, doi: 10.3847/0004-637X/820/1/38
  • Houde et al. (2009) Houde, M., Vaillancourt, J. E., Hildebrand, R. H., Chitsazzadeh, S., & Kirby, L. 2009, ApJ, 706, 1504, doi: 10.1088/0004-637X/706/2/1504
  • Hu & Lazarian (2023) Hu, Y., & Lazarian, A. 2023, MNRAS, 524, 2379, doi: 10.1093/mnras/stad1996
  • Hu et al. (2021) Hu, Y., Lazarian, A., & Stanimirović, S. 2021, ApJ, 912, 2, doi: 10.3847/1538-4357/abedb7
  • Hu et al. (2022) Hu, Y., Lazarian, A., & Wang, Q. D. 2022, MNRAS, 511, 829, doi: 10.1093/mnras/stac159
  • Hu et al. (2019) Hu, Y., Yuen, K. H., Lazarian, V., et al. 2019, Nature Astronomy, 3, 776, doi: 10.1038/s41550-019-0769-0
  • Hull & Zhang (2019) Hull, C. L. H., & Zhang, Q. 2019, Frontiers in Astronomy and Space Sciences, 6, 3, doi: 10.3389/fspas.2019.00003
  • Hunter (2007) Hunter, J. D. 2007, Computing in Science and Engineering, 9, 90, doi: 10.1109/MCSE.2007.55
  • Jackson et al. (2006) Jackson, J. M., Rathborne, J. M., Shah, R. Y., et al. 2006, ApJS, 163, 145, doi: 10.1086/500091
  • Jenness et al. (2013) Jenness, T., Chapin, E. L., Berry, D. S., et al. 2013, SMURF: SubMillimeter User Reduction Facility, Astrophysics Source Code Library, record ascl:1310.007. http://ascl.net/1310.007
  • Jiao et al. (2022) Jiao, S., Lin, Y., Shui, X., et al. 2022, Science China Physics, Mechanics, and Astronomy, 65, 299511, doi: 10.1007/s11433-021-1902-3
  • Juvela et al. (2018) Juvela, M., Guillet, V., Liu, T., et al. 2018, A&A, 620, A26, doi: 10.1051/0004-6361/201833245
  • Kauffmann et al. (2008) Kauffmann, J., Bertoldi, F., Bourke, T. L., Evans, N. J., I., & Lee, C. W. 2008, A&A, 487, 993, doi: 10.1051/0004-6361:200809481
  • Kauffmann et al. (2010) Kauffmann, J., Pillai, T., Shetty, R., Myers, P. C., & Goodman, A. A. 2010, ApJ, 716, 433, doi: 10.1088/0004-637X/716/1/433
  • Koch et al. (2012a) Koch, P. M., Tang, Y.-W., & Ho, P. T. P. 2012a, ApJ, 747, 79, doi: 10.1088/0004-637X/747/1/79
  • Koch et al. (2012b) —. 2012b, ApJ, 747, 80, doi: 10.1088/0004-637X/747/1/80
  • Koch et al. (2018) Koch, P. M., Tang, Y.-W., Ho, P. T. P., et al. 2018, ApJ, 855, 39, doi: 10.3847/1538-4357/aaa4c1
  • Koch et al. (2022) —. 2022, ApJ, 940, 89, doi: 10.3847/1538-4357/ac96e3
  • Kong (2019) Kong, S. 2019, ApJ, 873, 31, doi: 10.3847/1538-4357/aaffd5
  • Kong et al. (2019) Kong, S., Arce, H. G., Maureira, M. J., et al. 2019, ApJ, 874, 104, doi: 10.3847/1538-4357/ab07b9
  • Krumholz et al. (2005) Krumholz, M. R., McKee, C. F., & Klein, R. I. 2005, Nature, 438, 332, doi: 10.1038/nature04280
  • Lamarre et al. (2010) Lamarre, J. M., Puget, J. L., Ade, P. A. R., et al. 2010, A&A, 520, A9, doi: 10.1051/0004-6361/200912975
  • Larson (1981) Larson, R. B. 1981, MNRAS, 194, 809, doi: 10.1093/mnras/194.4.809
  • Lazarian (2007) Lazarian, A. 2007, J. Quant. Spec. Radiat. Transf., 106, 225, doi: 10.1016/j.jqsrt.2007.01.038
  • Lazarian & Hoang (2007) Lazarian, A., & Hoang, T. 2007, MNRAS, 378, 910, doi: 10.1111/j.1365-2966.2007.11817.x
  • Lazarian & Pogosyan (2006) Lazarian, A., & Pogosyan, D. 2006, ApJ, 652, 1348, doi: 10.1086/508012
  • Lazarian & Vishniac (1999) Lazarian, A., & Vishniac, E. T. 1999, ApJ, 517, 700, doi: 10.1086/307233
  • Lazarian & Yuen (2018) Lazarian, A., & Yuen, K. H. 2018, ApJ, 853, 96, doi: 10.3847/1538-4357/aaa241
  • Lazarian et al. (2018) Lazarian, A., Yuen, K. H., Ho, K. W., et al. 2018, ApJ, 865, 46, doi: 10.3847/1538-4357/aad7ff
  • Lazarian et al. (2022) Lazarian, A., Yuen, K. H., & Pogosyan, D. 2022, ApJ, 935, 77, doi: 10.3847/1538-4357/ac6877
  • Li (2021) Li, H.-B. 2021, Galaxies, 9, 41, doi: 10.3390/galaxies9020041
  • Li et al. (2009) Li, H.-b., Dowell, C. D., Goodman, A., Hildebrand, R., & Novak, G. 2009, ApJ, 704, 891, doi: 10.1088/0004-637X/704/2/891
  • Li et al. (2015a) Li, H.-B., Yuen, K. H., Otto, F., et al. 2015a, Nature, 520, 518, doi: 10.1038/nature14291
  • Li et al. (2018) Li, P. S., Klein, R. I., & McKee, C. F. 2018, MNRAS, 473, 4220, doi: 10.1093/mnras/stx2611
  • Li et al. (2022) Li, P. S., Lopez-Rodriguez, E., Ajeddig, H., et al. 2022, MNRAS, 510, 6085, doi: 10.1093/mnras/stab3448
  • Li et al. (2015b) Li, P. S., McKee, C. F., & Klein, R. I. 2015b, MNRAS, 452, 2500, doi: 10.1093/mnras/stv1437
  • Lin et al. (2016) Lin, Y., Liu, H. B., Li, D., et al. 2016, ApJ, 828, 32, doi: 10.3847/0004-637X/828/1/32
  • Lin et al. (2017) Lin, Y., Liu, H. B., Dale, J. E., et al. 2017, ApJ, 840, 22, doi: 10.3847/1538-4357/aa6c67
  • Liu et al. (2022a) Liu, J., Qiu, K., & Zhang, Q. 2022a, ApJ, 925, 30, doi: 10.3847/1538-4357/ac3911
  • Liu et al. (2021) Liu, J., Zhang, Q., Commerçon, B., et al. 2021, ApJ, 919, 79, doi: 10.3847/1538-4357/ac0cec
  • Liu et al. (2022b) Liu, J., Zhang, Q., & Qiu, K. 2022b, Frontiers in Astronomy and Space Sciences, 9, 943556, doi: 10.3389/fspas.2022.943556
  • Liu et al. (2020) Liu, J., Zhang, Q., Qiu, K., et al. 2020, ApJ, 895, 142, doi: 10.3847/1538-4357/ab9087
  • Liu et al. (2019) Liu, J., Qiu, K., Berry, D., et al. 2019, ApJ, 877, 43, doi: 10.3847/1538-4357/ab0958
  • Liu et al. (2023a) Liu, J., Zhang, Q., Koch, P. M., et al. 2023a, ApJ, 945, 160, doi: 10.3847/1538-4357/acb540
  • Liu et al. (2023b) Liu, J., Zhang, Q., Liu, H. B., et al. 2023b, ApJ, 949, 30, doi: 10.3847/1538-4357/acc4c0
  • Liu et al. (2018) Liu, T., Li, P. S., Juvela, M., et al. 2018, ApJ, 859, 151, doi: 10.3847/1538-4357/aac025
  • Mac Low & Klessen (2004) Mac Low, M.-M., & Klessen, R. S. 2004, Reviews of Modern Physics, 76, 125, doi: 10.1103/RevModPhys.76.125
  • Mairs et al. (2021) Mairs, S., Dempsey, J. T., Bell, G. S., et al. 2021, AJ, 162, 191, doi: 10.3847/1538-3881/ac18bf
  • McKee & Ostriker (2007) McKee, C. F., & Ostriker, E. C. 2007, ARA&A, 45, 565, doi: 10.1146/annurev.astro.45.051806.110602
  • McKee & Tan (2002) McKee, C. F., & Tan, J. C. 2002, Nature, 416, 59, doi: 10.1038/416059a
  • McMullin et al. (2007) McMullin, J. P., Waters, B., Schiebel, D., Young, W., & Golap, K. 2007, in Astronomical Society of the Pacific Conference Series, Vol. 376, Astronomical Data Analysis Software and Systems XVI, ed. R. A. Shaw, F. Hill, & D. J. Bell, 127
  • Mouschovias & Ciolek (1999) Mouschovias, T. C., & Ciolek, G. E. 1999, in NATO Advanced Study Institute (ASI) Series C, Vol. 540, The Origin of Stars and Planetary Systems, ed. C. J. Lada & N. D. Kylafis, 305
  • Mouschovias et al. (2006) Mouschovias, T. C., Tassis, K., & Kunz, M. W. 2006, ApJ, 646, 1043, doi: 10.1086/500125
  • Müller et al. (2001) Müller, H. S. P., Thorwirth, S., Roth, D. A., & Winnewisser, G. 2001, A&A, 370, L49, doi: 10.1051/0004-6361:20010367
  • Naghizadeh-Khouei & Clarke (1993) Naghizadeh-Khouei, J., & Clarke, D. 1993, A&A, 274, 968
  • Nakamura & Li (2008) Nakamura, F., & Li, Z.-Y. 2008, ApJ, 687, 354, doi: 10.1086/591641
  • Ngoc et al. (2023) Ngoc, N. B., Diep, P. N., Hoang, T., et al. 2023, ApJ, 953, 66, doi: 10.3847/1538-4357/acdb6e
  • Ossenkopf & Mac Low (2002) Ossenkopf, V., & Mac Low, M. M. 2002, A&A, 390, 307, doi: 10.1051/0004-6361:20020629
  • Ostriker et al. (2001) Ostriker, E. C., Stone, J. M., & Gammie, C. F. 2001, ApJ, 546, 980, doi: 10.1086/318290
  • Padoan et al. (2020) Padoan, P., Pan, L., Juvela, M., Haugbølle, T., & Nordlund, Å. 2020, ApJ, 900, 82, doi: 10.3847/1538-4357/abaa47
  • Pattle & Fissel (2019) Pattle, K., & Fissel, L. 2019, Frontiers in Astronomy and Space Sciences, 6, 15, doi: 10.3389/fspas.2019.00015
  • Pattle et al. (2023) Pattle, K., Fissel, L., Tahani, M., Liu, T., & Ntormousi, E. 2023, in Astronomical Society of the Pacific Conference Series, Vol. 534, Astronomical Society of the Pacific Conference Series, ed. S. Inutsuka, Y. Aikawa, T. Muto, K. Tomida, & M. Tamura, 193
  • Peretto et al. (2023) Peretto, N., Rigby, A. J., Louvet, F., et al. 2023, MNRAS, 525, 2935, doi: 10.1093/mnras/stad2453
  • Pillai et al. (2015) Pillai, T., Kauffmann, J., Tan, J. C., et al. 2015, ApJ, 799, 74, doi: 10.1088/0004-637X/799/1/74
  • Pillai et al. (2006) Pillai, T., Wyrowski, F., Carey, S. J., & Menten, K. M. 2006, A&A, 450, 569, doi: 10.1051/0004-6361:20054128
  • Planck Collaboration et al. (2014) Planck Collaboration, Abergel, A., Ade, P. A. R., et al. 2014, A&A, 571, A11, doi: 10.1051/0004-6361/201323195
  • Planck Collaboration et al. (2016) Planck Collaboration, Ade, P. A. R., Aghanim, N., et al. 2016, A&A, 586, A138, doi: 10.1051/0004-6361/201525896
  • Planck Collaboration et al. (2020) Planck Collaboration, Akrami, Y., Ashdown, M., et al. 2020, A&A, 641, A4, doi: 10.1051/0004-6361/201833881
  • Rathborne et al. (2006) Rathborne, J. M., Jackson, J. M., & Simon, R. 2006, ApJ, 641, 389, doi: 10.1086/500423
  • Remazeilles et al. (2011) Remazeilles, M., Delabrouille, J., & Cardoso, J.-F. 2011, MNRAS, 418, 467, doi: 10.1111/j.1365-2966.2011.19497.x
  • Sanhueza et al. (2012) Sanhueza, P., Jackson, J. M., Foster, J. B., et al. 2012, ApJ, 756, 60, doi: 10.1088/0004-637X/756/1/60
  • Sanhueza et al. (2019) Sanhueza, P., Contreras, Y., Wu, B., et al. 2019, ApJ, 886, 102, doi: 10.3847/1538-4357/ab45e9
  • Sanhueza et al. (2021) Sanhueza, P., Girart, J. M., Padovani, M., et al. 2021, ApJ, 915, L10, doi: 10.3847/2041-8213/ac081c
  • Savage & Jenkins (1972) Savage, B. D., & Jenkins, E. B. 1972, ApJ, 172, 491, doi: 10.1086/151369
  • Schöier et al. (2005) Schöier, F. L., van der Tak, F. F. S., van Dishoeck, E. F., & Black, J. H. 2005, A&A, 432, 369, doi: 10.1051/0004-6361:20041729
  • Simon et al. (2006) Simon, R., Rathborne, J. M., Shah, R. Y., Jackson, J. M., & Chambers, E. T. 2006, ApJ, 653, 1325, doi: 10.1086/508915
  • Skalidis & Tassis (2021) Skalidis, R., & Tassis, K. 2021, A&A, 647, A186, doi: 10.1051/0004-6361/202039779
  • Soam et al. (2019) Soam, A., Liu, T., Andersson, B. G., et al. 2019, ApJ, 883, 95, doi: 10.3847/1538-4357/ab39dd
  • Soler et al. (2013) Soler, J. D., Hennebelle, P., Martin, P. G., et al. 2013, ApJ, 774, 128, doi: 10.1088/0004-637X/774/2/128
  • Tan et al. (2014) Tan, J. C., Beltrán, M. T., Caselli, P., et al. 2014, in Protostars and Planets VI, ed. H. Beuther, R. S. Klessen, C. P. Dullemond, & T. Henning, 149–172, doi: 10.2458/azu_uapress_9780816531240-ch007
  • Tang et al. (2019) Tang, Y.-W., Koch, P. M., Peretto, N., et al. 2019, ApJ, 878, 10, doi: 10.3847/1538-4357/ab1484
  • Vaillancourt (2006) Vaillancourt, J. E. 2006, PASP, 118, 1340, doi: 10.1086/507472
  • Vázquez-Semadeni et al. (2019) Vázquez-Semadeni, E., Palau, A., Ballesteros-Paredes, J., Gómez, G. C., & Zamora-Avilés, M. 2019, MNRAS, 490, 3061, doi: 10.1093/mnras/stz2736
  • Wang (2018) Wang, K. 2018, Research Notes of the American Astronomical Society, 2, 52, doi: 10.3847/2515-5172/aacb29
  • Wang et al. (2012) Wang, K., Zhang, Q., Wu, Y., Li, H.-b., & Zhang, H. 2012, ApJ, 745, L30, doi: 10.1088/2041-8205/745/2/L30
  • Wang et al. (2008) Wang, Y., Zhang, Q., Pillai, T., Wyrowski, F., & Wu, Y. 2008, ApJ, 672, L33, doi: 10.1086/524949
  • Wang et al. (2006) Wang, Y., Zhang, Q., Rathborne, J. M., Jackson, J., & Wu, Y. 2006, ApJ, 651, L125, doi: 10.1086/508939
  • Wilson et al. (2013) Wilson, T. L., Rohlfs, K., & Hüttemeister, S. 2013, Tools of Radio Astronomy, doi: 10.1007/978-3-642-39950-3
  • Yuen & Lazarian (2017a) Yuen, K. H., & Lazarian, A. 2017a, ApJ, 837, L24, doi: 10.3847/2041-8213/aa6255
  • Yuen & Lazarian (2017b) —. 2017b, arXiv e-prints, arXiv:1703.03026. https://arxiv.longhoe.net/abs/1703.03026
  • Zhang et al. (2015) Zhang, Q., Wang, K., Lu, X., & Jiménez-Serra, I. 2015, ApJ, 804, 141, doi: 10.1088/0004-637X/804/2/141
  • Zhang et al. (2009) Zhang, Q., Wang, Y., Pillai, T., & Rathborne, J. 2009, ApJ, 696, 268, doi: 10.1088/0004-637X/696/1/268
  • Zhang et al. (2014) Zhang, Q., Qiu, K., Girart, J. M., et al. 2014, ApJ, 792, 116, doi: 10.1088/0004-637X/792/2/116
  • Zweibel (1990) Zweibel, E. G. 1990, ApJ, 362, 545, doi: 10.1086/169291