{CJK*}

UTF8gbsn

\volnopage

Vol.0 (20xx) No.0, 000–000

11institutetext: Kavli Institute for Astronomy and Astrophysics, Peking University, Beijing 100871, People’s Republic of China; [email protected], [email protected]
22institutetext: Department of Astronomy, School of Physics, Peking University, Beijing, 100871, People’s Republic of China
33institutetext: Shanghai Astronomical Observatory, Chinese Academy of Sciences, 80 Nandan Road, Shanghai 200030, People’s Republic of China; [email protected]
44institutetext: Chinese Academy of Sciences South America Center for Astronomy, National Astronomical Observatories, Chinese Academy of Sciences, Beijing, 100101, People’s Republic of China;
55institutetext: Departamento de Astronomía, Universidad de Chile, Las Condes, 7591245 Santiago, Chile;
66institutetext: Jet Propulsion Laboratory, California Institute of Technology, 4800 Oak Grove Drive, Pasadena CA 91109, USA;
77institutetext: Center for Astrophysics |||| Harvard & Smithsonian, 60 Garden Street, Cambridge, MA 02138, USA;
88institutetext: National Astronomical Observatory of Japan, National Institutes of Natural Sciences, 2-21-1 Osawa, Mitaka, Tokyo 181-8588, Japan;
99institutetext: Astronomical Science Program, The Graduate University for Advanced Studies, SOKENDAI, 2-21-1 Osawa, Mitaka, Tokyo 181-8588, Japan;
1010institutetext: School of Physics and Astronomy, Yunnan University, Kunming 650091, People’s Republic of China;
1111institutetext: Yunnan Observatories, Chinese Academy of Sciences, 396 Yangfangwang, Guandu District, Kunming, 650216, People’s Republic of China;
1212institutetext: Department of Physics, University of Helsinki, PO Box 64, FI-00014 Helsinki, Finland;
1313institutetext: Indian Institute of Space Science and Technology, Thiruvananthapuram 695 547, Kerala, India;
1414institutetext: Max Planck Institute for Astronomy, Königstuhl 17, D-69117 Heidelberg, Germany;
1515institutetext: Department of Astronomy, Graduate School of Science, The University of Tokyo, 7-3-1 Hongo, Bunkyo-ku, Tokyo 113-0033, Japan;
1616institutetext: Max-Planck-Institut für Radioastronomie, Auf dem Hügel 69, 53121 Bonn, Germany;
1717institutetext: Departamento de Astronom´ıa, Universidad de Concepcion, Casilla 160-C, Concepci ´ on, Chile;
1818institutetext: Department of Astronomy, The University of Texas at Austin, Texas 78712-1205, USA;
1919institutetext: Korea Astronomy and Space Science Institute, 776 Daedeokdae-ro, Yuseong-gu, Daejeon 34055, Republic of Korea;
2020institutetext: Institute of Astronomy and Astrophysics, Academia Sinica, Roosevelt Road, Taipei 10617, Taiwan (R.O.C);
2121institutetext: South-Western Institute for Astronomy Research, Yunnan University, Kunming, People’s Republic of China;
2222institutetext: Astronomical Science Program, The Graduate University for Advanced Studies, SOKENDAI, 2-21-1 Osawa, Mitaka, Tokyo 181-8588, Japan;
2323institutetext: Korea Astronomy and Space Science Institute, 776 Daedeok-daero, Yuseong-gu, Daejeon 34055, Republic of Korea;
2424institutetext: University of Science and Technology, 217 Gajeong-ro, Yuseong-gu, Daejeon 34113, Republic of Korea;
2525institutetext: National Astronomical Observatories, Chinese Academy of Sciences, Beijing 100101, People’s Republic of China;
2626institutetext: Department of Mathematical Sciences, University of South Africa, Cnr Christian de Wet Rd and Pioneer Avenue, Florida Park, 1709, Roodepoort, South Africa;
2727institutetext: Centre for Space Research, Physics Department, North-West University, Potchefstroom 2520, South Africa;
2828institutetext: Department of Physics and Astronomy, Faculty of Physical Sciences, University of Nigeria, Carver Building, 1 University Road, Nsukka 410001, Nigeria;
2929institutetext: Institute of Physics and Astronomy, Eötvös Lorànd University, Pázmány Péter sétány 1/A, H-1117 Budapest, Hungary;
3030institutetext: University of Debrecen, Faculty of Science and Technology, Egyetem tér 1, H-4032 Debrecen, Hungary;
\vs\noReceived 2023 month day; accepted 2023 month day

The ALMA-QUARKS Survey: II. the ACA 1.3 mm continuum source catalog and the assembly of dense gas in massive star-forming clumps

Fengwei Xu (许峰玮) 1122    Ke Wang 11    Tie Liu 33    Lei Zhu 44    Guido Garay 55    Xunchuan Liu 33    Paul Goldsmith 66    Qizhou Zhang 77    Patricio Sanhueza 8899    Shengli Qin 1010    Jinhua He 11114455    Mika Juvela 1212    Anandmayee Tej 1313    Hongli Liu 1010    Shanghuo Li 1414    Kaho Morii 151588    Siju Zhang 11    Jianwen Zhou 1616    Amelia Stutz 1717    Neal J. Evans 1818    Kim Kee-Tae 1919    Shengyuan Liu 2020    Diego Mardones 55    Guangxing Li 2121    Leonardo Bronfman 55    Ken’ichi Tatematsu 882222    Chang Won Lee 23232424    Xing Lu 33    Xiaofeng Mai 33    Sihan Jiao 2525    James O. Chibueze 262627272828    Keyun Su 11    L. Viktor Tóth 29293030
Abstract

Leveraging the high resolution, sensitivity, and wide frequency coverage of the Atacama Large Millimeter/submillimeter Array (ALMA), the QUARKS survey, standing for ‘Querying Underlying mechanisms of massive star formation with ALMA-Resolved gas Kinematics and Structures’, is observing 139 massive star-forming clumps at ALMA Band 6 (λsimilar-to𝜆absent\lambda\simitalic_λ ∼ 1.3 mm). This paper introduces the Atacama Compact Array (ACA) 7-m data of the QUARKS survey, describing the ACA observations and data reduction. Combining multi-wavelength data, we provide the first edition of QUARKS atlas, offering insights into the multiscale and multiphase interstellar medium (ISM) in high-mass star formation. The ACA 1.3 mm catalog includes 207 continuum sources that are called ACA sources. Their gas kinetic temperatures are estimated using three formaldehyde transitions with a non-LTE radiation transfer model, and the mass and density are derived from a dust emission model. The ACA sources are massive (16–84 percentile values of 6–160 Msubscript𝑀direct-productM_{\odot}italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT), gravity-dominated (MR1.1proportional-to𝑀superscript𝑅1.1M\propto R^{1.1}italic_M ∝ italic_R start_POSTSUPERSCRIPT 1.1 end_POSTSUPERSCRIPT) fragments within massive clumps, with supersonic turbulence (>11\mathcal{M}>1caligraphic_M > 1) and embedded star-forming protoclusters. We find a linear correlation between the masses of the fragments and the massive clumps, with a ratio of 6% between the two. When considering the fragments as representative of dense gas, the ratio indicates a dense gas fraction (DGF) of 6%, although with a wide scatter ranging from 1% to 10%. If we consider the QUARKS massive clumps to be what is observed at various scales, then the size-independent DGF indicates a self-similar fragmentation or collapsing mode in protocluster formation. With the ACA data over four orders of magnitude of luminosity-to-mass ratio (L/M𝐿𝑀L/Mitalic_L / italic_M), we find that the DGF increases significantly with L/M𝐿𝑀L/Mitalic_L / italic_M, which indicates clump evolutionary stage. We observed a limited fragmentation at the subclump scale, which can be explained by dynamic global collapse process.

keywords:
stars: formation — stars: kinematics and dynamics — ISM: clouds — ISM: molecules

1 Introduction

High-mass stars (M>8𝑀8M>8italic_M > 8Msubscript𝑀direct-productM_{\odot}italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT), as the principal sources of UV radiation and heavy elements, play a major role in the evolution of galaxies. However, the properties and evolution of massive clumps hosting the precursors of massive stars currently forming in our Galaxy are still poorly known (e.g., Zinnecker & Yorke 2007; Tan et al. 2014; Pineda et al. 2023). Fragmentation of massive clumps into dense cores, where star formation ultimately occurs, is a critical step in the mass assembly process that gives rise to stars and clusters, as highlighted by Motte et al. (2018). Investigating how dense gas is concentrated and structured within these massive clumps serves as an intermediate step in understanding this intricate process.

Recent studies, such as Peretto et al. (2020), have conducted sub-millimeter continuum surveys of infrared dark clouds, revealing that the evolution of massive compact sources in mass-versus-temperature diagrams is better explained by an accretion scenario where cores gain mass while simultaneously collapsing to form protostars. Furthermore, the findings from Rigby et al. (2021) provide evidence for the mass growth of clumps, suggesting that similar mass accumulation processes may occur on a broader range of physical scales, which is further verified in several multiscale case studies (e.g., Neupane et al. 2020; Xu et al. 2023a; Yang et al. 2023). On the simulation side, the mass growth of a ‘core’ is believed to be the result of the collapse of the surrounding parsec-scale mass reservoir called a ‘clump’, hence the accretion scenario described above is referred to as ‘clump-fed’ (Wang et al. 2010).

The ‘clump-fed’ scenario suggests that there must be a link between the properties of a clump and the fragments that form within it. In the case of clump fragmentation, Lin et al. (2019) find a correlation between the mass of the clump and the mass of its most massive fragment of massive clumps from “the APEX Telescope Large Area Survey of the Galaxy” (ATLASGAL; Schuller et al. 2009). Besides, Barnes et al. (2021) also find that more massive, and more turbulent clouds make more 0.1similar-toabsent0.1\sim 0.1∼ 0.1 pc scale cores. Anderson et al. (2021) collected a sample of massive clumps with a wide range of evolutionary stages, and suggested time-dependent correlation between clump and core mass especially in hub-filament systems. The relation between the formation of dense cores and the properties of clumps such as turbulence is also discussed (e.g., Xu et al. 2021, 2024a). On a smaller scale, studies such as Palau et al. (2014, 2021) find correlations between the level of fragmentation within massive dense cores (<0.1absent0.1<0.1< 0.1 pc) and their average volume density, aligning with the expectations of the Jeans instability (Sanhueza et al. 2019; Morii et al. 2024). Furthermore, using ALMA with a resolution of approximately 0.02 pc, Xu et al. (2024b) identified a sublinear correlation between the mass of the clump and the mass of its most massive core, in a sample of 11 massive protoclusters that show clump-scale infall motion. For comparison, no such correlation was found in a sample of 39 massive early-stage clumps from Morii et al. (2023). This suggests that the mass correlation between the clumps and the cores gradually builds up over the evolution of massive clumps.

Despite previous great advances, our understanding of the formation process of massive stars remains unclear and divided due to observational difficulties. On the one hand, considering their large distances (a few kpc) and high dust extinction, studies of massive clumps need high-resolution interferometric observations to resolve their internal gas structures and kinematics (Wang 2015; Motte et al. 2018; Lin 2021). On the other hand, obtaining robust and definitive conclusions regarding the accretion history of high-mass stars requires a larger statistical sample. This, in turn, calls for rapid survey capabilities with adequate sensitivity. The Atacama Large Millimeter/submillimeter Array (ALMA), with both high resolution and high sensitivity, offers a unique valuable opportunity to investigate the hierarchical structures in massive star-forming regions in great detail. Therefore, we performed a 1.3-mm ALMA survey called “Querying Underlying mechanisms of massive star formation with ALMA-Resolved gas Kinematics and Structures” (QUARKS; PIs: Lei Zhu, Guido Garay and Tie Liu; Project ID: 2021.1.00095.S). The details of the survey description and the showcase of data combinations can be found in Liu et al. (2023b).

In this paper, we focus mainly on the Atacama Compact Array (ACA) 7-m continuum and line data sets. With relatively little free-free emission contamination and a maximum recoverable scale as large as 27similar-toabsent27\sim 27∼ 27\arcsec, ACA 1.3 mm continuum data are useful for tracking dense molecular gas within massive clumps. With an angular resolution of 5similar-toabsent5\sim 5∼ 5\arcsec, equivalent to 0.07 pc at a typical distance of 3 kpc within the QUARKS sample, the ACA data provide a global view of massive protostars or protoclusters therein. We first introduce the ACA observations and data imaging in Section 2, and then provide the first edition of the QUARKS atlas in Section 3. Section 4 constructs the ACA 1.3 mm continuum source catalog including physical parameters. In Section 5.1, we discuss the physical nature of ACA sources and find that they are condensed gas fragments within massive clumps. In Section 5.2, we find a mass correlation between clumps and their fragments. In Section 5.3, we discuss the size-invariant and time-variant dense gas fraction. In Section 5.4, we discuss limited fragmentation at the subclump scale. In Section 6, we present a brief summary.

2 QUARKS ACA Data

2.1 Observing Strategy

QUARKS acts as a follow-up 1.3-mm survey of ALMA Three-millimeter Observations of Massive Star-forming regions (ATOMS; Liu et al. 2020), and aims at studying even smaller structures within the 3-mm cores or core clusters within massive star-forming clumps. To ensure uniformity of the sample and solid detection at 1.3 mm, we exclude: 1) two low-mass clumps (<15absent15<15< 15 Msubscript𝑀direct-productM_{\odot}italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT); 2) four sources dominated by extended Hii regions with angular sizes larger than the primary beam at Band 6; 3) one source without continuum emission detection by ATOMS. As a result, 139 ATOMS massive clumps are selected as the QUARKS sample, hereafter the QUARKS clumps. A total of 156 ALMA 1.3 mm pointings were performed because some 3 mm emission show elongated or extended morphologies and need two mosaicked pointings (dual-pointing mosaicked field). The on-source time of each pointing was about 5 minutes.

The QUARKS ACA observations are separated into 15 scheduling blocks (SBs), hereafter called groups for short. The group ID and the number of fields therein are listed in columns (1)–(2) of Table 1. To finish the observing queue, 1–3 execution blocks (EBs) were performed on different observing dates, which are listed in column (3). Fields in the same group have the proximity of sky coordinates and the same EBs, so they share similar minimum and maximum baselines (BL), angular resolution (AR), and maximum recoverable scale (MRS), which are listed in columns (4)–(6). Variations in AR and MRS are mostly the result of different configuration of the array and source elevation. Phase and flux calibrators are listed in column (7), while bandpass and flux calibrators are listed in column (8).

The observations utilized ALMA Band 6 receivers in dual-polarization mode, with the correlator frequencies configured into four spectral windows (SPWs 1–4). The four SPWs were designed with center frequencies at approximately 217.92 GHz, 220.32 GHz, 231.37 GHz, and 233.52 GHz, each with a bandwidth of 2 GHz and 4096 channels. This setup was chosen to cover a wide range of commonly used tracers representing different environments and excitation conditions. The targeted lines included, but were not limited to: 1) outflow tracers (e.g., CO, 13CO, SiO, SO, H2CO); 2) cold gas tracer (N2D+); 3) dense core and filament tracers (e.g., C18O, HC3N); 4) hot core tracers (e.g., CH3OH, C2H5CN, NH2CHO, CH3CN); 5) ionized-gas/Hii-region tracer (Hα30subscript𝛼30{}_{30}\alphastart_FLOATSUBSCRIPT 30 end_FLOATSUBSCRIPT italic_α). The basics of the main targeted lines are summarized in Table 2 of Liu et al. (2023b).

Table 1: QUARKS ACA Observation and Imaging Result Logs
Group ID nfieldsubscript𝑛fieldn_{\rm field}italic_n start_POSTSUBSCRIPT roman_field end_POSTSUBSCRIPTa Obs. Date Min./Max. BL AR MRS Calibrators cont. rms
(m/m) (\arcsec) (\arcsec) Phase Bandpass/Flux (mJy beam-1)
(1) (2) (3) (4) (5) (6) (7) (8) (9)
1 6 2021-11-06 8.9/45.0 4.9 20.0 J0922-3959 J1058+0133 10
2 3 2021-10-22 8.9/44.7 5.1 28.3 J1047-6217 J1058+0133 6
2022-05-21 8.9/45.0 4.9 28.7 J1047-6217 J1058+0133
3 7 2023-01-02 8.9/48.9 5.0 28.3 J1424-6807 J1427-4206 9
4 12 2023-01-13 8.9/48.9 4.8 28.3 J1337-6509 J1427-4206 7
2023-01-16 8.9/48.0 5.0 28.3 J1337-6509 J1427-4206
5 24 2023-04-09 8.9/48.0 4.5 21.1 J1604-4441 J1427-4206 10
2023-04-13 8.9/48.0 4.9 27.1 J1617-5848 J1427-4206
2023-04-18 8.9/48.0 4.9 27.1 J1617-5848 J1427-4206
6 5 2023-01-01 8.9/48.9 5.3 28.3 J1524-5903 J1427-4206 18
7 13 2023-05-17 8.9/48.9 4.8 27.1 J1733-3722 J1924-2914 9
2023-05-20 8.9/48.9 4.8 27.1 J1733-3722 J1924-2914
8 3 2023-01-14 8.9/48.0 4.9 21.1 J1924+1540 J2232+1143 43
9 3 2023-01-22 8.9/48.0 5.0 28.3 J1744-3116 J1427-4206 22
10 8 2023-03-04 9.1/45.0 4.9 21.1 J1851+0035 J1924-2914 5
2023-04-16 8.9/48.0 4.9 27.1 J1851+0035 J1924-2914
11 7 2023-04-08 8.9/48.0 4.5 21.1 J1717-3342 J1427-4206 28
12 18 2023-04-25 8.9/48.0 4.9 27.1 J1604-4441 J1427-4206 5
2023-05-01 8.9/48.0 4.9 27.1 J1604-4441 J1427-4206
13 11 2023-03-04 9.1/45.0 4.9 21.1 J1851+0035 J1924-2914 8
2023-04-09 8.9/48.0 4.5 21.1 J1851+0035 J1924-2914
14 6 2023-04-21 8.9/48.0 4.9 27.1 J1820-2528 J1924-2914 17
15 13 2023-04-17 8.9/48.0 4.9 27.1 J1832-2039 J1924-2914 12
2023-04-18 8.9/48.0 4.9 27.1 J1832-2039 J1924-2914

The QUARKS ACA observations are separated into 15 scheduling blocks which are called groups for short. The group ID, the number of targets therein, and the dates of observation are listed in columns (1)–(3). The minimum and maximum baselines (BL) of the configuration array are listed in column (4). Group-averaged angular resolution (AR) and maximum recoverable scale (MRS) are listed in columns (5)–(6). The phase calibrator(s) are listed in column (7) while the bandpass and flux calibrators are listed in column (8). The aggregated continuum imaging rms and spectral line rms per channel are listed in column (9).

(a)Including both single-pointing and dual-pointing fields, therefore 139 in total.

2.2 Data Calibration and Imaging

QUARKS ACA data were acquired during the ALMA Cycle 8 and 9 observations. The data were routinely calibrated using the ALMA pipelines 111https://almascience.nrao.edu/processing/science-pipeline of Common Astronomy Software Applications (CASA; McMullin et al. 2007; CASA Team et al. 2022) in the corresponding versions of 6.2.1 and 6.4.1. The frequency tunings of the correlator are four wide bands, each with 2 GHz separated by 4096 channels. The edge channels of each spectral window (2×128similar-toabsent2128\sim 2\times 128∼ 2 × 128) were flagged in the first version of data release due to the high temperature of the system noise and therefore the high level of noise.

Line emission channels were flagged to obtain the continuum and spectral lines simultaneously. Liu et al. (2023b) identified all the transitions of strong lines within the four SPWs by matching the reduced data cubes of the ALMA pipeline and the laboratory databases for the spectral lines (CDMS; Müller et al. 2001). The QUARKS team generated a model spectrum as a mask for line emission channels. For each source, the model spectrum was shifted and expanded with a width of 50 km s-1 to ensure clean channels free from multiple velocity components and spectral line wings. With this method the line emission channels were flagged and the line-free channels were subtracted from the four spectral windows in the Fourier space using the task uvcontsub with linear fitting (polynomial order of 1). The continuum and line cube imaging processes were performed by the task tclean in CASA 6.5.6, with briggs robust weighting of 0.5. In the cleaning process, the images/cubes were automatically masked by auto-multithresh algorithm whose input parameters are recommended by the official guides 222casaguide:automask for the ACA data. At the beginning of each minor cycle, the cleaning mask was updated on the basis of the current residual image. The algorithm uses multiple thresholds based on the noise and sidelobe levels in the residual image to determine the cleaning mask. Within the cleaning mask, we set the stopping noisethreshold to be 5σ𝜎\sigmaitalic_σ and sidelobethreshold to be 1.25σ1.25𝜎1.25\sigma1.25 italic_σ in the dirty image/cube. However, the cleaning algorithm diverges when some fields have relatively strong emission or side lobes at the edge. For these fields, we performed a manual mask to further improve the imaging fidelity. To recover the potential large-scale structures in the spectral lines and mitigate artifacts produced by extended emission, we applied the multiscale deconvolver (Cornwell 2008) to the cleaning process of line cubes, with scales of [0,5,15]. We uniformly set the image size of 108×108108108108\times 108108 × 108 pix2 and cell size of 1\arcsec, to fully cover the 17 mosaic fields with dual pointings. The primary beam correction was performed with pblimit = 0.2.

The noise rms of the cleaned continuum image is tabulated in column (9) of Table 1. The rms levels have large variations between groups because some sources are too strong and the continuum sensitivities are limited by the dynamic range. Self-calibration can improve the sensitivity, but will not contribute much to the science goals in this paper. Therefore, self-calibration has not been performed on our released ACA data.

3 QUARKS Atlas

With the inclusion of QUARKS as a high-resolution submillimeter dataset, we are now equipped to construct a comprehensive data atlas for the QUARKS sample, offering insights into the multiscale and multiphase interstellar medium (ISM) in high-mass star formation.

The mid infrared (MIR) Spitzer facility was equipped with the InfraRed Array Camera (IRAC) instrument that provided images at 3.6, 4.5, 5.8, and 8.0 μ𝜇\muitalic_μm simultaneously. Here we combine the Galactic Legacy Infrared Mid-Plane Survey Extraordinaire (GLIMPSE) data at 3.6, 4.5, and 8.0 μ𝜇\muitalic_μm into the pseudo color map in the left panel of Figure 1. The three bands can provide information about young stellar objects (YSOs), shocked molecular gas in protostellar outflows, and hot dust or polycyclic aromatic hydrocarbon (PAH) emission, respectively. The angular resolution of the images is smoothed to a uniform value of 2.02.02.02.0\arcsec. Some of our source positions have no available Spitzer data, and for these we use ALLWISE data at 3.4, 4.6, and 22.0 μ𝜇\muitalic_μm (Wright et al. 2010; Mainzer et al. 2011) instead, which have an angular resolution of 6.\arcsec.\!\!\arcsec.1, 6.\arcsec.\!\!\arcsec.4, 6.\arcsec.\!\!\arcsec.5, and 12.\arcsec.\!\!\arcsec.0, respectively.

The submillimeter emission traces the cold and dense gas well. ATLASGAL explores the inner Galactic plane at submillimeter wavelengths (870similar-toabsent870\sim 870∼ 870μ𝜇\muitalic_μm) with a beam size of 19.\arcsec.\!\!\arcsec.2 (Schuller et al. 2009). We use the ATLASGAL data to trace the large scale cold and dense gas which indicates the star formation region, in the white contours on the left panel. If no ATLASGAL data are available, far infrared (FIR) Herschel data from “the Herschel Infrared Galactic Plane Survey” (HiGAL; Traficante et al. 2011) at 500 μ𝜇\muitalic_μm is used instead, with a nominal beam size of 34.\arcsec.\!\!\arcsec.5 (Traficante et al. 2011).

The MeerKAT Galactic Plane Survey 1.28 GHz data (MGPS; Padmanabh et al. 2023; Goedhart et al. 2023) provides essential radio information with a resolution of 8\arcsecsimilar-toabsent8\arcsec\sim 8\arcsec∼ 8. It is overlaid with yellow contours to indicate the ionized gas from ultra-compact Hii (UCHii) and hyper-compact Hii (HCHii) regions as well as radio jets.

On a smaller scale, the ATOMS 12m+ACA combined 3 mm continuum data can trace both the dust emission from cold dense cores and/or the free-free emission from UCHii and HCHii regions (Liu et al. 2021). The potential contamination from free-free emission can be estimated using centimeter wavelength data (Avison et al. 2015; Olguin et al. 2022; Xu et al. 2023a, and forthcoming ATCA project).

In Figure 1, the red circle indicates the field of view (80similar-toabsent80\sim 80∼ 80\arcsec) of the ATOMS continuum data, zooming in on the substructures of 2\arcsecsimilar-toabsent2\arcsec\sim 2\arcsec∼ 2 in one of the QUARKS massive clump I13291-6229, as shown in the middle panel. The 3 mm continuum emission of I13291-6229 exhibits a filamentary morphology from the southeast to the northwest. Dense cores identified in Liu et al. (2021) are marked with red ellipses and ID numbers. As a follow-up, the QUARKS 1.3 mm ACA observations are pointed towards the 3 mm continuum emission region, as marked by the white dashed circles where the 7-m primary beam response is 0.2. In the right panel, the QUARKS ACA data are shown in the background color map and four 1.3 mm sources are identified, where three are identified as solid detection with SNR>9absent9>9> 9 in red and one with SNR<9absent9<9< 9 in yellow.

Refer to caption
Figure 1: QUARKS multi-band atlas of representative source I13291-6229. Left panel: the background is the Spitzer 3.6/4.5/8 μ𝜇\muitalic_μm pseudo color map, overlaid with Herschel 500 μ𝜇\muitalic_μm (white contours) and MeerKAT Galactic Plane Survey (MGPS) 1.28 GHz data (yellow contours). The red circle indicates the field of view (80similar-toabsent80\sim 80∼ 80\arcsec) of the combined ATOMS 12m + ACA 3 mm continuum data. Middle panel: the background is the ATOMS combined 3 mm continuum data, linearly scaled from 9σ9𝜎-9\sigma- 9 italic_σ to 9σ9𝜎9\sigma9 italic_σ and logarithmically scaled from 9σ9𝜎9\sigma9 italic_σ to peak intensity. The source IDs are in order from North to South, and the nomenclature follows “#Field_ATOMS#ID”. The green dashed circle(s) indicate the QUARKS pointing(s), with size of 7-m primary beam response of 0.2. The ATOMS beam size is shown on the bottom left. Right panel: the background is the QUARKS ACA 1.3 mm continuum data, linearly scaled from 3σ3𝜎-3\sigma- 3 italic_σ to peak intensity. The continuum sources are shown as red ellipses (SNR>9absent9>9> 9) and yellow ellipses (SNR<9absent9<9< 9). The source IDs are in order from North to South and the nomenclature follows “#Field_ACA#ID”. The QUARKS beam size is shown in the bottom left. The scale bars in three panels are shown on the bottom right.

The QUARKS survey updates the clump distances using the H13CO+ lines of the ATOMS survey and the latest model for the rotation curve of the Milky Way (Reid et al. 2019), as listed in Table A. of Liu et al. (2023b). At the bottom right of Figure 1, both the angular and physical scale bars are shown, with updated clump distance. The complete QUARKS data atlases are presented in the supplementary material.

4 Results

4.1 ACA 1.3 mm Continuum Source Catalog

We adopted an automatic source extraction algorithm SExtractor 333https://sextractor.readthedocs.io/en/latest/Introduction.html. (Bertin & Arnouts 1996) on the ACA 1.3 mm continuum emission maps to extract the 1.3 mm continuum sources. The advantages of SExtractor in our case are: 1) to subtract the background diffuse emission and rms noise automatically; 2) to support local rms noise input as pixelwise thresholds; 3) to deblend the potentially overlapped sources in crowded fields. The details for the algorithm input parameters are described in Appendix A.

As a result, a total of 207 ACA 1.3 mm continuum sources are extracted from the 139 massive star-forming clumps and the fundamental measurements are summarized in Table 2. At least one and at most five sources have been detected in one clump. The field name is listed in column (1) and the continuum source ID which is given in order from North to South is listed in column (2). Hereafter, the format of ACA 1.3 mm continuum source obeys “#Field_ACA#ID”. The ICRS coordinates of the sources are listed in columns (3)–(4). The FWHM of the major and minor axes and the position angle are listed in columns (5)–(7). The integrated flux and peak intensity, which are corrected by primary beam (see details in Section A), are listed in columns (8)–(9). The signal-to-noise (SNR) ratio, defined as the peak intensity over the local rms, is listed in column (10).

We note that the maximum observed angular size (convolved with the beam) of ACA sources is 14\arcsecsimilar-toabsent14\arcsec\sim 14\arcsec∼ 14, which is only half of the maximum recoverable scale in most cases (see Table 1). We thus ignore the effects of missing flux in the following analyses.

Table 2: Basic Measurements of ACA 1.3 mm Continuum Sources .
Field ID Equatorial Coordinates θmaj×θminsubscript𝜃majsubscript𝜃min\theta_{\rm maj}\times\theta_{\rm min}italic_θ start_POSTSUBSCRIPT roman_maj end_POSTSUBSCRIPT × italic_θ start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT PA Fintsubscript𝐹intF_{\rm int}italic_F start_POSTSUBSCRIPT roman_int end_POSTSUBSCRIPT Ipeaksubscript𝐼peakI_{\rm peak}italic_I start_POSTSUBSCRIPT roman_peak end_POSTSUBSCRIPT SNR
ACA RA (ICRS) DEC (ICRS) (arcsec) (arcsec) (deg) (Jy) (Jy beam-1)
(1) (2) (3) (4) (5) (6) (7) (8) (9) (10)
I08303-4303 1 08:32:08.7 -43:13:46.2 10.1 7.0 128.6 0.811 0.347 31.8
I08448-4343 1 08:46:35.1 -43:54:22.6 6.8 5.1 64.9 0.143 0.108 11.4
I08448-4343 2 08:46:34.7 -43:54:33.5 7.6 4.1 156.6 0.087 0.076 9.5
I08448-4343 3 08:46:33.4 -43:54:36.5 7.5 4.6 5.9 0.171 0.126 16.5
I08448-4343 4 08:46:32.5 -43:54:37.1 12.2 4.4 5.2 0.175 0.114 16.5
I08470-4243 1 08:48:47.8 -42:54:26.4 10.0 6.0 74.6 0.989 0.623 77.2
I09002-4732 1 09:01:54.3 -47:44:09.8 7.2 5.8 2.6 2.675 1.720 31.2
I09018-4816 1 09:03:33.3 -48:28:01.0 12.1 7.0 96.9 1.403 0.617 40.5
I09094-4803 1 09:11:08.6 -48:15:44.1 5.3 5.0 164.7 0.063 0.056 14.3
I09094-4803 2 09:11:08.3 -48:15:53.3 6.6 5.2 81.4 0.093 0.073 22.3

The field name and the continuum source ID are listed in columns (1)–(2). The ICRS coordinates of the barycenter are listed in columns (3)–(4). The FWHM of the major and minor axes (θmajsubscript𝜃maj\theta_{\rm maj}italic_θ start_POSTSUBSCRIPT roman_maj end_POSTSUBSCRIPT and θminsubscript𝜃min\theta_{\rm min}italic_θ start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT), and the position angle (PA) of sources are listed in columns (5)–(7). The integrated flux Fintsubscript𝐹intF_{\rm int}italic_F start_POSTSUBSCRIPT roman_int end_POSTSUBSCRIPT and the peak intensity Ipeaksubscript𝐼peakI_{\rm peak}italic_I start_POSTSUBSCRIPT roman_peak end_POSTSUBSCRIPT are listed in columns (8)–(9). The signal-to-noise ratio (SNR) is listed in column (10). The table is available in its entirety in machine-readable form.

4.2 Temperature Estimation from Formaldehyde

Formaldehyde (H2CO) is a suitable spectroscopic tool to derive the kinetic temperature of the molecular gas (Ao et al. 2013). For instance, observations throughout the Central Molecular Zone (CMZ; Ginsburg et al. 2016), show that the H2CO emission is spatially widespread and correlated with dust emission. Formaldehyde has two isomeric species, ortho-H2CO (o-H2CO) and para-H2CO (p-H2CO). Kahane et al. (1984) reported that p-H2CO is 1–3 times less abundant than o-H2CO in three low-mass star forming regions. Mangum & Wootten (1993) reported ortho-to-para ratio of 1.5–3 in the Orion-KL high-mass star-forming region. Therefore, temperature estimation from p-H2CO transition lines minimizes the uncertainties from a high optical depth. For example, Tang et al. (2017) find that the kinetic temperatures derived from p-H2CO show a good agreement with those of the dust in the Large Magellanic Cloud (LMC), suggesting that the dust and p-H2CO molecules are probing the same (or similar) gas component. More importantly, Tang et al. (2021) mapped two massive star forming regions in the Large Magellanic Cloud (LMC) using ALMA at a resolution of 0.4 pc and find a consistency in the spatial distribution of the dense gas traced by p-H2CO with that of the 1.3 mm dust. Above all, these studies provide a practical foundation for using the p-H2CO molecule to estimate the dust temperature of the ACA sources.

The QUARKS frequency tunings in SPW1 are designed to cover the p-H2CO transition triplet, i.e., JKA,Kc=30,320,2subscript𝐽subscriptKAsubscriptKcsubscript303subscript202J_{\rm K_{A},K_{c}}=3_{0,3}\rightarrow 2_{0,2}italic_J start_POSTSUBSCRIPT roman_K start_POSTSUBSCRIPT roman_A end_POSTSUBSCRIPT , roman_K start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT end_POSTSUBSCRIPT = 3 start_POSTSUBSCRIPT 0 , 3 end_POSTSUBSCRIPT → 2 start_POSTSUBSCRIPT 0 , 2 end_POSTSUBSCRIPT (303subscript3033_{03}3 start_POSTSUBSCRIPT 03 end_POSTSUBSCRIPT202subscript2022_{02}2 start_POSTSUBSCRIPT 02 end_POSTSUBSCRIPT) at 218.22219 GHz, 32,222,1subscript322subscript2213_{2,2}\rightarrow 2_{2,1}3 start_POSTSUBSCRIPT 2 , 2 end_POSTSUBSCRIPT → 2 start_POSTSUBSCRIPT 2 , 1 end_POSTSUBSCRIPT (322subscript3223_{22}3 start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT221subscript2212_{21}2 start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT) at 218.47563 GHz, and 32,122,0subscript321subscript2203_{2,1}\rightarrow 2_{2,0}3 start_POSTSUBSCRIPT 2 , 1 end_POSTSUBSCRIPT → 2 start_POSTSUBSCRIPT 2 , 0 end_POSTSUBSCRIPT (321subscript3213_{21}3 start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT220subscript2202_{20}2 start_POSTSUBSCRIPT 20 end_POSTSUBSCRIPT) at 218.76007 GHz. Their upper state energies are 21.0 K, 68.09 K, and 68.11 K, respectively. We extracted spectra from 218.1 to 281.9 GHz in SPW1 to fully cover the frequency range of the H2CO triplet, and perform a five-parameter model using the formaldehyde official model444https://pyspeckit.readthedocs.io/en/latest/formaldehyde_model.html by pyspeckit (Ginsburg & Mirocha 2011; Ginsburg et al. 2022) with a non-LTE RADEX model. The details of the fitting model and results are summarized in Appendix B. The derived kinetic temperatures have values between 24 to 180 K with median of 72 K, which are listed in column (3) of Table 3.

Refer to caption
Figure 2: Dust temperature at the scales of clump (Tdust,clumpsubscript𝑇dustclumpT_{\rm dust,clump}italic_T start_POSTSUBSCRIPT roman_dust , roman_clump end_POSTSUBSCRIPT) and embedded ACA fragment (Tdust,fragsubscript𝑇dustfragT_{\rm dust,frag}italic_T start_POSTSUBSCRIPT roman_dust , roman_frag end_POSTSUBSCRIPT). Tdust,fragsubscript𝑇dustfragT_{\rm dust,frag}italic_T start_POSTSUBSCRIPT roman_dust , roman_frag end_POSTSUBSCRIPT is derived from the non-LTE H2CO modeling based on the assumption that dust temperature is identical to gas temperature. Tdust,clumpsubscript𝑇dustclumpT_{\rm dust,clump}italic_T start_POSTSUBSCRIPT roman_dust , roman_clump end_POSTSUBSCRIPT is derived from infrared dust emission SED fitting. The “isothermal line” with Tdust,frag=Tdust,clumpsubscript𝑇dustfragsubscript𝑇dustclumpT_{\rm dust,frag}=T_{\rm dust,clump}italic_T start_POSTSUBSCRIPT roman_dust , roman_frag end_POSTSUBSCRIPT = italic_T start_POSTSUBSCRIPT roman_dust , roman_clump end_POSTSUBSCRIPT is shown in orange. The uncertainty of Tdust,fragsubscript𝑇dustfragT_{\rm dust,frag}italic_T start_POSTSUBSCRIPT roman_dust , roman_frag end_POSTSUBSCRIPT is given by model fitting and the relative uncertainty of Tdust,clumpsubscript𝑇dustclumpT_{\rm dust,clump}italic_T start_POSTSUBSCRIPT roman_dust , roman_clump end_POSTSUBSCRIPT is set to be 10% for all sources.

A mixture of gas and dust having density over 104.5superscript104.510^{4.5}10 start_POSTSUPERSCRIPT 4.5 end_POSTSUPERSCRIPT cm-3 due to collisional process (Goldsmith 2001) yields Tdustsubscript𝑇dustT_{\rm dust}italic_T start_POSTSUBSCRIPT roman_dust end_POSTSUBSCRIPT equal to Tkinsubscript𝑇kinT_{\rm kin}italic_T start_POSTSUBSCRIPT roman_kin end_POSTSUBSCRIPT, which holds for most of the ACA sources (refer to Section 5.1.2). Therefore, in this work, we assume that the dust temperature of ACA source/fragment (Tdust,fragsubscript𝑇dustfragT_{\rm dust,frag}italic_T start_POSTSUBSCRIPT roman_dust , roman_frag end_POSTSUBSCRIPT) equals to the kinetic temperature derived from H2CO triplet line fitting (Tkin,H2COsubscript𝑇kinsubscriptH2COT_{\rm kin,H_{2}CO}italic_T start_POSTSUBSCRIPT roman_kin , roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT roman_CO end_POSTSUBSCRIPT).

Figure 2 illustrates the relationship between fragment-scale dust temperature Tdust,fragsubscript𝑇dustfragT_{\rm dust,frag}italic_T start_POSTSUBSCRIPT roman_dust , roman_frag end_POSTSUBSCRIPT and the clump-averaged dust temperature Tdust,clumpsubscript𝑇dustclumpT_{\rm dust,clump}italic_T start_POSTSUBSCRIPT roman_dust , roman_clump end_POSTSUBSCRIPT determined through infrared-to-submillimeter SED fitting (see method in Urquhart et al. 2018). All the data points lie above the “isothermal line” (Tkin,H2CO=Tdust,clumpsubscript𝑇kinsubscriptH2COsubscript𝑇dustclumpT_{\rm kin,H_{2}CO}=T_{\rm dust,clump}italic_T start_POSTSUBSCRIPT roman_kin , roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT roman_CO end_POSTSUBSCRIPT = italic_T start_POSTSUBSCRIPT roman_dust , roman_clump end_POSTSUBSCRIPT), shown in orange color. This suggests an outward negative temperature gradient, with the inner dense gas traced by H2CO being warmer than the surrounding low-density gas and warmer than gas on average. The temperature gradient serves as an indirect observational evidence of the idea that the ACA sources are active star-forming regions. We note that seven sources with failed fitting (six with a strong self-absorption line profile and one with a weak detection) are assigned with the same value as Tdust,clumpsubscript𝑇dustclumpT_{\rm dust,clump}italic_T start_POSTSUBSCRIPT roman_dust , roman_clump end_POSTSUBSCRIPT, therefore on the orange line in Figure 2.

4.3 Physical Parameters of Sources

Assuming that all the emission comes from dust in a single component with Tdustsubscript𝑇dustT_{\rm dust}italic_T start_POSTSUBSCRIPT roman_dust end_POSTSUBSCRIPT and that the dust emission is optically thin, the masses of the sources can be calculated using

Msource=FintD2κνBν(Tdust),subscript𝑀sourcesubscript𝐹intsuperscript𝐷2subscript𝜅𝜈subscript𝐵𝜈subscript𝑇dustM_{\rm source}=\mathcal{R}\frac{F_{\rm int}D^{2}}{\kappa_{\nu}B_{\nu}(T_{\rm dust% })},italic_M start_POSTSUBSCRIPT roman_source end_POSTSUBSCRIPT = caligraphic_R divide start_ARG italic_F start_POSTSUBSCRIPT roman_int 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 start_POSTSUBSCRIPT roman_dust end_POSTSUBSCRIPT ) end_ARG , (1)

where Fintsubscript𝐹intF_{\rm int}italic_F start_POSTSUBSCRIPT roman_int end_POSTSUBSCRIPT is the measured integrated flux of dust emission, \mathcal{R}caligraphic_R is the gas-to-dust mass ratio (assumed to be 100), D𝐷Ditalic_D is the clump distance, κνsubscript𝜅𝜈\kappa_{\nu}italic_κ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT is the dust opacity per gram of dust, and Bν(Tdust)subscript𝐵𝜈subscript𝑇dustB_{\nu}(T_{\rm dust})italic_B start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ( italic_T start_POSTSUBSCRIPT roman_dust end_POSTSUBSCRIPT ) is the Planck function at a given dust temperature Tdustsubscript𝑇dustT_{\rm dust}italic_T start_POSTSUBSCRIPT roman_dust end_POSTSUBSCRIPT. In our case, κνsubscript𝜅𝜈\kappa_{\nu}italic_κ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT is assumed to be 1 cm2 g-1 at ν230similar-to𝜈230\nu\sim 230italic_ν ∼ 230 GHz which is interpolated from the given table in Ossenkopf & Henning (1994), assuming grains with thin ice mantles and the size distribution given by Mathis et al. (1977) and a typical gas density of 106superscript10610^{6}10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT cm-3 in our sample. Substituting Tdustsubscript𝑇dustT_{\rm dust}italic_T start_POSTSUBSCRIPT roman_dust end_POSTSUBSCRIPT assumed above to Equation 1, the ACA source masses are then calculated and listed in the column (4). The major sources of uncertainty in the mass calculation come from the gas-to-dust ratio and the dust opacity. We adopt the uncertainties derived by Sanhueza et al. (2017) of 28% for the gas-to-dust ratio and of 23% for the dust opacity, contributing to the 36similar-toabsent36\sim 36∼ 36% uncertainty of the specific dust opacity. The uncertainty of Fintsubscript𝐹intF_{\rm int}italic_F start_POSTSUBSCRIPT roman_int end_POSTSUBSCRIPT from flux calibration (assumed to be 10%; Yun et al. 2022 555ALMA Memo 211) and the uncertainty of distance (assumed to be 20%) are included. Monte Carlo methods are adopted for uncertainty estimation and 1σ1𝜎1\sigma1 italic_σ confidence intervals are given.

ACA sources are characterized by 2D Gaussian-like ellipses with the FWHM of the major and minor axes (θmajsubscript𝜃maj\theta_{\rm maj}italic_θ start_POSTSUBSCRIPT roman_maj end_POSTSUBSCRIPT and θminsubscript𝜃min\theta_{\rm min}italic_θ start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT), and position angle (PA) listed in columns (5)–(6) of Table 2. Following Rosolowsky et al. (2010) and Contreras et al. (2013); Urquhart et al. (2014), the source angular size can be calculated as the geometric mean of the deconvolved major and minor axes.

θdec=η[(σmaj2σbm2)(σmin2σbm2)]1/4,subscript𝜃dec𝜂superscriptdelimited-[]subscriptsuperscript𝜎2majsubscriptsuperscript𝜎2bmsubscriptsuperscript𝜎2minsubscriptsuperscript𝜎2bm14\theta_{\rm dec}=\eta\left[\left(\sigma^{2}_{\rm maj}-\sigma^{2}_{\rm bm}% \right)\left(\sigma^{2}_{\rm min}-\sigma^{2}_{\rm bm}\right)\right]^{1/4},italic_θ start_POSTSUBSCRIPT roman_dec end_POSTSUBSCRIPT = italic_η [ ( italic_σ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_maj end_POSTSUBSCRIPT - italic_σ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_bm end_POSTSUBSCRIPT ) ( italic_σ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT - italic_σ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_bm end_POSTSUBSCRIPT ) ] start_POSTSUPERSCRIPT 1 / 4 end_POSTSUPERSCRIPT , (2)

where σmajsubscript𝜎maj\sigma_{\rm maj}italic_σ start_POSTSUBSCRIPT roman_maj end_POSTSUBSCRIPT and σminsubscript𝜎min\sigma_{\rm min}italic_σ start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT are calculated from θmaj/8ln2subscript𝜃maj82\theta_{\rm maj}/\sqrt{8\ln 2}italic_θ start_POSTSUBSCRIPT roman_maj end_POSTSUBSCRIPT / square-root start_ARG 8 roman_ln 2 end_ARG and θmin/8ln2subscript𝜃min82\theta_{\rm min}/\sqrt{8\ln 2}italic_θ start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT / square-root start_ARG 8 roman_ln 2 end_ARG respectively. σbmsubscript𝜎bm\sigma_{\rm bm}italic_σ start_POSTSUBSCRIPT roman_bm end_POSTSUBSCRIPT is the averaged dispersion size of the beam (i.e., θbmajθbmin/(8ln2)subscript𝜃bmajsubscript𝜃bmin82\sqrt{\theta_{\rm bmaj}\theta_{\rm bmin}/(8\ln 2)}square-root start_ARG italic_θ start_POSTSUBSCRIPT roman_bmaj end_POSTSUBSCRIPT italic_θ start_POSTSUBSCRIPT roman_bmin end_POSTSUBSCRIPT / ( 8 roman_ln 2 ) end_ARG where θbmjsubscript𝜃bmj\theta_{\rm bmj}italic_θ start_POSTSUBSCRIPT roman_bmj end_POSTSUBSCRIPT and θbminsubscript𝜃bmin\theta_{\rm bmin}italic_θ start_POSTSUBSCRIPT roman_bmin end_POSTSUBSCRIPT are the FWHM of the major and minor axis of the beam). η𝜂\etaitalic_η is a factor that relates the size of the emission distribution dispersion to the determined angular radius of the object. η=2.4𝜂2.4\eta=2.4italic_η = 2.4, the median value derived for a range of models consisting of a spherical emissivity distribution (Rosolowsky et al. 2010), is adopted here. Therefore, the physical size can be calculated directly using Rdec=θdec×Dsubscript𝑅decsubscript𝜃dec𝐷R_{\rm dec}=\theta_{\rm dec}\times Ditalic_R start_POSTSUBSCRIPT roman_dec end_POSTSUBSCRIPT = italic_θ start_POSTSUBSCRIPT roman_dec end_POSTSUBSCRIPT × italic_D, as shown in column (5) of Table 3.

The source peak column density is estimated from

NH2peak=IpeakΩμH2mHκνBν(Tdust),subscriptsuperscript𝑁peaksubscriptH2subscript𝐼peakΩsubscript𝜇subscriptH2subscript𝑚Hsubscript𝜅𝜈subscript𝐵𝜈subscript𝑇dustN^{\rm peak}_{\scriptscriptstyle\rm H_{2}}=\mathcal{R}\frac{I_{\rm peak}}{% \Omega\mu_{\scriptscriptstyle\rm H_{2}}m_{\scriptscriptstyle\rm H}\kappa_{\nu}% B_{\nu}(T_{\rm dust})},italic_N start_POSTSUPERSCRIPT roman_peak end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT = caligraphic_R divide start_ARG italic_I start_POSTSUBSCRIPT roman_peak end_POSTSUBSCRIPT end_ARG start_ARG roman_Ω italic_μ start_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_m start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT italic_κ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT italic_B start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ( italic_T start_POSTSUBSCRIPT roman_dust end_POSTSUBSCRIPT ) end_ARG , (3)

where Ipeaksubscript𝐼peakI_{\rm peak}italic_I start_POSTSUBSCRIPT roman_peak end_POSTSUBSCRIPT is the measured peak flux of source within the beam solid angle ΩΩ\Omegaroman_Ω.

The surface density averaged by the source can be calculated by Σ=Msource/(πRdec2)Σsubscript𝑀source𝜋subscriptsuperscript𝑅2dec\Sigma=M_{\rm source}/(\pi R^{2}_{\rm dec})roman_Σ = italic_M start_POSTSUBSCRIPT roman_source end_POSTSUBSCRIPT / ( italic_π italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_dec end_POSTSUBSCRIPT ). The source-averaged number density, nH2subscript𝑛subscriptH2n_{\rm H_{2}}italic_n start_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT, is then calculated by assuming a spherical source,

nH2=Msource(4/3)πμH2mHRdec3,subscript𝑛subscriptH2subscript𝑀source43𝜋subscript𝜇subscriptH2subscript𝑚Hsuperscriptsubscript𝑅dec3n_{\scriptscriptstyle\rm H_{2}}=\frac{M_{\rm source}}{(4/3)\pi\mu_{% \scriptscriptstyle\rm H_{2}}m_{\scriptscriptstyle\rm H}R_{\rm dec}^{3}},italic_n start_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT = divide start_ARG italic_M start_POSTSUBSCRIPT roman_source end_POSTSUBSCRIPT end_ARG start_ARG ( 4 / 3 ) italic_π italic_μ start_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_m start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT roman_dec end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG , (4)

where μH2subscript𝜇subscriptH2\mu_{\scriptscriptstyle\rm H_{2}}italic_μ start_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT is the molecular weight per hydrogen molecule and mHsubscript𝑚Hm_{\rm H}italic_m start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT is the mass of a hydrogen atom. Throughout the paper, we adopt the molecular weight per hydrogen molecule μH2=2.81subscript𝜇subscriptH22.81\mu_{\scriptscriptstyle\rm H_{2}}=2.81italic_μ start_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT = 2.81 (Evans et al. 2022). The calculated peak column density and the average surface density and volume densities of the source are given in columns (6)–(8) of Table 3.

The velocity dispersion contributed by the thermal motion of H2CO molecules is given by

σth,H2CO=kBTkinmH2CO,subscript𝜎thsubscriptH2COsubscript𝑘𝐵subscript𝑇kinsubscript𝑚subscriptH2CO\sigma_{\scriptscriptstyle\rm th,H_{2}CO}=\sqrt{\frac{k_{B}T_{\rm kin}}{m_{\rm H% _{2}CO}}},italic_σ start_POSTSUBSCRIPT roman_th , roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT roman_CO end_POSTSUBSCRIPT = square-root start_ARG divide start_ARG italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT roman_kin end_POSTSUBSCRIPT end_ARG start_ARG italic_m start_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT roman_CO end_POSTSUBSCRIPT end_ARG end_ARG , (5)

where Tkinsubscript𝑇kinT_{\rm kin}italic_T start_POSTSUBSCRIPT roman_kin end_POSTSUBSCRIPT is kinetic temperature derived in Section 4.2 and mH2COsubscript𝑚subscriptH2COm_{\rm H_{2}CO}italic_m start_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT roman_CO end_POSTSUBSCRIPT is the molecular weight 30 times mHsubscript𝑚𝐻m_{H}italic_m start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT. With σH2COsubscript𝜎subscriptH2CO\sigma_{\rm H_{2}CO}italic_σ start_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT roman_CO end_POSTSUBSCRIPT deduced from the observed velocity dispersion σobssubscript𝜎obs\sigma_{\rm obs}italic_σ start_POSTSUBSCRIPT roman_obs end_POSTSUBSCRIPT, the non-thermal velocity dispersion is derived as

σnt=σobs2Δchan2/(22ln2)2σth,H2CO2,subscript𝜎ntsubscriptsuperscript𝜎2obssuperscriptsubscriptΔchan2superscript2222subscriptsuperscript𝜎2thsubscriptH2CO\sigma_{\rm nt}=\sqrt{\sigma^{2}_{\rm obs}-\Delta_{\rm chan}^{2}/(2\sqrt{2\ln 2% })^{2}-\sigma^{2}_{\rm th,H_{2}CO}},italic_σ start_POSTSUBSCRIPT roman_nt end_POSTSUBSCRIPT = square-root start_ARG italic_σ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_obs end_POSTSUBSCRIPT - roman_Δ start_POSTSUBSCRIPT roman_chan end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ( 2 square-root start_ARG 2 roman_ln 2 end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_σ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_th , roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT roman_CO end_POSTSUBSCRIPT end_ARG , (6)

where Δchan=1.34subscriptΔchan1.34\Delta_{\rm chan}=1.34roman_Δ start_POSTSUBSCRIPT roman_chan end_POSTSUBSCRIPT = 1.34 km s-1 is the channel width. We check that 30 sources with line width >7absent7>7> 7 km s-1 are strongly influenced by outflow wings (see Appendix B), and set an averaged observed velocity dispersion of 1.661.661.661.66 km s-1.

The gas sound speed (cssubscript𝑐𝑠c_{s}italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT) or thermal velocity dispersion (σthsubscript𝜎th\sigma_{\rm th}italic_σ start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT) is given by,

cs=σth=kBTkinμpmH,subscript𝑐𝑠subscript𝜎thsubscript𝑘𝐵subscript𝑇kinsubscript𝜇𝑝subscript𝑚Hc_{s}=\sigma_{\rm th}=\sqrt{\frac{k_{B}T_{\rm kin}}{\mu_{p}m_{% \scriptscriptstyle\rm H}}},italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = italic_σ start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT = square-root start_ARG divide start_ARG italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT roman_kin end_POSTSUBSCRIPT end_ARG start_ARG italic_μ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT italic_m start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT end_ARG end_ARG , (7)

where μp=2.37subscript𝜇𝑝2.37\mu_{p}=2.37italic_μ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = 2.37 for a mean molecular weight per free particle (Kauffmann et al. 2008). The three-dimensional (3D) Mach number is defined as,

=3σntcs.3subscript𝜎ntsubscript𝑐𝑠\mathcal{M}=\frac{\sqrt{3}\sigma_{\rm nt}}{c_{s}}.caligraphic_M = divide start_ARG square-root start_ARG 3 end_ARG italic_σ start_POSTSUBSCRIPT roman_nt end_POSTSUBSCRIPT end_ARG start_ARG italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_ARG . (8)

The sound speed, non-thermal velocity dispersion and Mach number are listed in columns (9)–(11) of Table 3.

Table 3: Physical Parameters of ACA 1.3 mm Continuum Sources
Field ID Tkinsubscript𝑇kinT_{\rm kin}italic_T start_POSTSUBSCRIPT roman_kin end_POSTSUBSCRIPT Msourcesubscript𝑀sourceM_{\rm source}italic_M start_POSTSUBSCRIPT roman_source end_POSTSUBSCRIPT Rdecsubscript𝑅decR_{\rm dec}italic_R start_POSTSUBSCRIPT roman_dec end_POSTSUBSCRIPT NH2peaksubscriptsuperscript𝑁peaksubscriptH2N^{\rm peak}_{\rm H_{2}}italic_N start_POSTSUPERSCRIPT roman_peak end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT ΣΣ\Sigmaroman_Σ nH2subscript𝑛subscriptH2n_{\rm H_{2}}italic_n start_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT cssubscript𝑐𝑠c_{s}italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT σntsubscript𝜎nt\sigma_{\rm nt}italic_σ start_POSTSUBSCRIPT roman_nt end_POSTSUBSCRIPT \mathcal{M}caligraphic_M
ACA (K) (Msubscript𝑀direct-productM_{\odot}italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT) (pc) (cm-2) (g cm-2) (cm-3) (km s-1) (km s-1)
(1) (2) (3) (4) (5) (6) (7) (8) (9) (10) (11)
I13291-6229 1 32.8(6.3) 2.7(1.4) 1.7(0.7)×1022absentsuperscript1022\times 10^{22}× 10 start_POSTSUPERSCRIPT 22 end_POSTSUPERSCRIPT 0.34 1 3.0
I13291-6229 2 74.8(1.4) 8.1(3.5) 0.054 3.3(1.2)×1022absentsuperscript1022\times 10^{22}× 10 start_POSTSUPERSCRIPT 22 end_POSTSUPERSCRIPT 0.18(0.08) 1.8(0.8)×105absentsuperscript105\times 10^{5}× 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT 0.51 1.4 2.7
I13291-6229 3 56.6(0.7) 3.5(1.5) 0.032 1.8(0.7)×1022absentsuperscript1022\times 10^{22}× 10 start_POSTSUPERSCRIPT 22 end_POSTSUPERSCRIPT 0.23(0.10) 3.7(1.6)×105absentsuperscript105\times 10^{5}× 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT 0.44 0.93 2.1
I13291-6229 4 66.6(1.8) 2.3(1.0) 1.1(0.4)×1022absentsuperscript1022\times 10^{22}× 10 start_POSTSUPERSCRIPT 22 end_POSTSUPERSCRIPT 0.48
I13291-6249 1 87.4(3.4) 281.3(122.0) 0.3 8.9(3.3)×1022absentsuperscript1022\times 10^{22}× 10 start_POSTSUPERSCRIPT 22 end_POSTSUPERSCRIPT 0.21(0.09) 3.6(1.6)×104absentsuperscript104\times 10^{4}× 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT 0.55 1.9 3.5
I13295-6152 1 44.5(0.6) 26.3(11.3) 0.12 5.4(2.0)×1022absentsuperscript1022\times 10^{22}× 10 start_POSTSUPERSCRIPT 22 end_POSTSUPERSCRIPT 0.13(0.05) 5.6(2.4)×104absentsuperscript104\times 10^{4}× 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT 0.39 0.96 2.5
I13471-6120 1 78.0(0.9) 119.7(51.6) 0.048 2.2(0.8)×1023absentsuperscript1023\times 10^{23}× 10 start_POSTSUPERSCRIPT 23 end_POSTSUPERSCRIPT 3.51(1.51) 3.8(1.6)×106absentsuperscript106\times 10^{6}× 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT 0.52 1.5 2.8
I13484-6100 1 73.0(14.5) 119.6(61.4) 0.1 9.9(3.9)×1022absentsuperscript1022\times 10^{22}× 10 start_POSTSUPERSCRIPT 22 end_POSTSUPERSCRIPT 0.74(0.38) 3.7(1.9)×105absentsuperscript105\times 10^{5}× 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT 0.5 2.5 5.0
I14013-6105 1 98.0(2.5) 83.1(35.9) 0.11 1.2(0.5)×1023absentsuperscript1023\times 10^{23}× 10 start_POSTSUPERSCRIPT 23 end_POSTSUPERSCRIPT 0.47(0.20) 2.2(1.0)×105absentsuperscript105\times 10^{5}× 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT 0.58 1.5 2.6
I14050-6056 1 120.0(6.9) 18.4(8.0) 0.11 3.5(1.3)×1022absentsuperscript1022\times 10^{22}× 10 start_POSTSUPERSCRIPT 22 end_POSTSUPERSCRIPT 0.10(0.04) 4.9(2.1)×104absentsuperscript104\times 10^{4}× 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT 0.65 1.5 2.3

Field name and the continuum source ID are listed in columns (1)–(2). Kinetic temperature is listed in column (3). Source mass (Msourcesubscript𝑀sourceM_{\rm source}italic_M start_POSTSUBSCRIPT roman_source end_POSTSUBSCRIPT), deconvolved size (Rdecsubscript𝑅decR_{\rm dec}italic_R start_POSTSUBSCRIPT roman_dec end_POSTSUBSCRIPT), peak column density (NH2peaksubscriptsuperscript𝑁peaksubscriptH2N^{\rm peak}_{\rm H_{2}}italic_N start_POSTSUPERSCRIPT roman_peak end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT), surface density (ΣΣ\Sigmaroman_Σ), and volume density (nH2subscript𝑛subscriptH2n_{\rm H_{2}}italic_n start_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT) are listed in columns (4)–(8). The uncertainties of these parameters (except for Rdecsubscript𝑅decR_{\rm dec}italic_R start_POSTSUBSCRIPT roman_dec end_POSTSUBSCRIPT) are included in parentheses. Unresolved sources have ‘–’ in columns (5), (7) and (8). Sound speed (cssubscript𝑐𝑠c_{s}italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT), non-thermal velocity dispersion (σntsubscript𝜎nt\sigma_{\rm nt}italic_σ start_POSTSUBSCRIPT roman_nt end_POSTSUBSCRIPT), and Mach number (\mathcal{M}caligraphic_M) are listed in columns (9)–(11). If H2CO fitting fails, then mark with ‘–’ in columns (10)–(11). Only a part of the table is shown and the complete table is available in its entirety in machine-readable form.

5 Discussion

5.1 Nature of ACA Source

Refer to caption
Figure 3: Histograms of (a) kinetic temperature Tkinsubscript𝑇kinT_{\rm kin}italic_T start_POSTSUBSCRIPT roman_kin end_POSTSUBSCRIPT, (b) source mass Msourcesubscript𝑀sourceM_{\rm source}italic_M start_POSTSUBSCRIPT roman_source end_POSTSUBSCRIPT, (c) surface density, (d) Mach number \mathcal{M}caligraphic_M, (e) source size Rdecsubscript𝑅decR_{\rm dec}italic_R start_POSTSUBSCRIPT roman_dec end_POSTSUBSCRIPT, and (f) volume density nH2subscript𝑛subscriptH2n_{\rm H_{2}}italic_n start_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT.

5.1.1 Massive star-forming regions with supersonic turbulence

Six parameters from Table 3, including kinetic temperature (Tkinsubscript𝑇kinT_{\rm kin}italic_T start_POSTSUBSCRIPT roman_kin end_POSTSUBSCRIPT), source mass (Msourcesubscript𝑀sourceM_{\rm source}italic_M start_POSTSUBSCRIPT roman_source end_POSTSUBSCRIPT), source size (Rdecsubscript𝑅decR_{\rm dec}italic_R start_POSTSUBSCRIPT roman_dec end_POSTSUBSCRIPT), volume density (nH2subscript𝑛subscriptH2n_{\rm H_{2}}italic_n start_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT), surface density (ΣΣ\Sigmaroman_Σ) and Mach number (\mathcal{M}caligraphic_M) are shown in histogram form in Figure 3.

In panel (a), Tkinsubscript𝑇kinT_{\rm kin}italic_T start_POSTSUBSCRIPT roman_kin end_POSTSUBSCRIPT ranges from 20similar-toabsent20\sim 20∼ 20 to 180similar-toabsent180\sim 180∼ 180 K, with a mean and median of similar-to\sim68 K. More than 97% sources have a temperature larger than 30 K, indicating their protostellar nature. With a higher resolution of 2\arcsec, Qin et al. (2022) identified 60 hot molecular cores with gas temperature >100absent100>100> 100 K, as the heating sources of massive clumps. As shown in Figure 2, the temperatures of ACA sources are all larger than those of clumps, indicating that the ACA sources are heating their parent massive clumps. Theoretically, a massive protostellar embryo heats and eventually ionizes the gas of its surrounding envelope, creating an Hii region that develops by expanding within the cloud (Motte et al. 2018). Therefore, the embedded heating sources are expected, indicating ACA sources should be heating their parent massive clumps.

Panel (b) of Figure 3 shows the mass distribution, with 16 and 84 percentile values of 6 and 160 Msubscript𝑀direct-productM_{\odot}italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT. Assuming monolithic collapse and that only a single star forms, 120 ACA sources (58%) are massive enough to form a massive star (>8absent8>8> 8Msubscript𝑀direct-productM_{\odot}italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT) with a core-to-star efficiency 30similar-toabsent30\sim 30∼ 30% (Matzner & McKee 2000; Federrath & Klessen 2012). However, the ACA sources have fragmented and are forming protoclusters (see Section 5.1.3). If assuming that a cluster will form inside the ACA sources and using the empirical relation from star clusters given by Larson (2003),

(mmaxM)=1.2(MclusterM)0.45,subscript𝑚maxsubscript𝑀direct-product1.2superscriptsubscript𝑀clustersubscript𝑀direct-product0.45\left(\frac{m_{\rm max}}{M_{\odot}}\right)=1.2\left(\frac{M_{\rm cluster}}{M_{% \odot}}\right)^{0.45},( divide start_ARG italic_m start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT end_ARG start_ARG italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT end_ARG ) = 1.2 ( divide start_ARG italic_M start_POSTSUBSCRIPT roman_cluster end_POSTSUBSCRIPT end_ARG start_ARG italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT 0.45 end_POSTSUPERSCRIPT , (9)

where mmaxsubscript𝑚maxm_{\rm max}italic_m start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT and Mclustersubscript𝑀clusterM_{\rm cluster}italic_M start_POSTSUBSCRIPT roman_cluster end_POSTSUBSCRIPT are the maximum mass and the total mass of the stellar cluster, we find that 72 out of 207 sources can form massive stars if only mass in situ participates in star formation. From panel (c), we observe that nearly all the sources possess surface densities (ΣΣ\Sigmaroman_Σ) exceeding 0.05 g cm-2, which aligns with the empirical threshold for high-mass star formation as suggested by Urquhart et al. (2014). A total of 35 ACA sources exceed the more stringent surface density threshold of 1 g cm-2 proposed by Krumholz & McKee (2008). But it’s essential to bear in mind that these surface density thresholds can be scale-dependent. For example, if massive clumps have a density profile of ρR2proportional-to𝜌superscript𝑅2\rho\propto R^{-2}italic_ρ ∝ italic_R start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT, then the enclosed mass scales with MRproportional-to𝑀𝑅M\propto Ritalic_M ∝ italic_R, resulting in ΣR1proportional-toΣsuperscript𝑅1\Sigma\propto R^{-1}roman_Σ ∝ italic_R start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT. If we adjust for this relation, the surface density threshold for massive star formation should be approximately ten times higher at the scale of ACA sources. But in a turbulent-dominated clump, one argues that cores have a column density comparable to that of the clump as a whole, with only Σcore/Σcl=1.22subscriptΣcoresubscriptΣcl1.22\Sigma_{\rm core}/\Sigma_{\rm cl}=1.22roman_Σ start_POSTSUBSCRIPT roman_core end_POSTSUBSCRIPT / roman_Σ start_POSTSUBSCRIPT roman_cl end_POSTSUBSCRIPT = 1.22.

Core-scale (0.1 pc and n>105𝑛superscript105n>10^{5}italic_n > 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT cm-3) infall motions have been statistically studied (Wu & Evans 2003; Wu et al. 2007; Contreras et al. 2018; Xu et al. 2023b) and filamentary accretion flows are resolved in high-resolution observations (Peretto et al. 2013; Liu et al. 2016; Yuan et al. 2018; Lu et al. 2018; Chen et al. 2019; Sanhueza et al. 2021; Redaelli et al. 2022; Xu et al. 2023a; Yang et al. 2023). Hence, the identified ACA sources are likely to continue to accumulate mass throughout their evolution, achieving further growth of the core mass and enhancement of the surface density (Liu et al. 2023a; Xu et al. 2024b). Using H13CO+ (1–0), Zhou et al. (2022) identified 68 hub-filament systems with clear velocity gradients in the ATOMS survey. In the context of massive cluster formation, these hub-filament structures play a crucial role in supporting gas accretion towards dense cores where massive stars form.

In panel (d), the Mach numbers \mathcal{M}caligraphic_M at the scale of the ACA sources are mostly greater than 2. We note that those sources with strong line wings are excluded when analyzing the line widths because of the widening effects of the H2CO outflows as reported in (Izumi et al. 2023). Assuming that turbulence dominates the non-thermal motion, the \mathcal{M}caligraphic_M values suggest the prevalence of supersonic turbulence, which aligns with what has been found earlier in several cases (e.g., Zhang et al. 2009; Wang et al. 2014). As a result, the supersonic turbulence can suppress thermal Jeans fragmentation (1similar-toabsent1\sim 1∼ 1Msubscript𝑀direct-productM_{\odot}italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT; Sanhueza et al. 2019) and enhances mass accretion onto central massive protostars, which is proposed in the model by McKee & Tan (2003).

5.1.2 Fragments with self-similar gravitational collapse

Refer to caption
Figure 4: Mass versus radius (M𝑀Mitalic_MR𝑅Ritalic_R) diagram. The QUARKS clumps and the ACA fragments are shown with orange stars and blue pentagons, respectively. Two linear regressions are applied to the QUARKS clumps and the ACA fragments in logarithmic space. The corresponding scaling relations MR1.80proportional-to𝑀superscript𝑅1.80M\propto R^{1.80}italic_M ∝ italic_R start_POSTSUPERSCRIPT 1.80 end_POSTSUPERSCRIPT and MR1.11proportional-to𝑀superscript𝑅1.11M\propto R^{1.11}italic_M ∝ italic_R start_POSTSUPERSCRIPT 1.11 end_POSTSUPERSCRIPT are shown with orange and blue solid lines. The Spearman correlation coefficient and 1σ𝜎\sigmaitalic_σ scatter are shown in the lower right corner of the figure. The orange and blue shaded regions correspond to the turbulence-dominated (MR2proportional-to𝑀superscript𝑅2M\propto R^{2}italic_M ∝ italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT) and gravity-dominated (MRproportional-to𝑀𝑅M\propto Ritalic_M ∝ italic_R) regimes with 2σ2𝜎2\sigma2 italic_σ data scatters in linear regression fittings.

According to panels (e) and (f) in Figure 3, the ACA sources exhibit sizes ranging from 0.02 to 0.4 pc, volume densities exceeding 104superscript10410^{4}10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT cm-3, with 107 of them surpassing 105superscript10510^{5}10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT cm-3. The median volume density of the ACA sources is 1.6×1051.6superscript1051.6\times 10^{5}1.6 × 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT cm-3, which is 20 times larger than that of the clumps. This suggests that these ACA sources are the condensed gas fragments at a subclump scale. Therefore, the ACA sources serve as sub-clump structures and are also called (ACA) fragments in the following discussion.

We collect the QUARKS clump mass Mclumpsubscript𝑀clumpM_{\rm clump}italic_M start_POSTSUBSCRIPT roman_clump end_POSTSUBSCRIPT and radius Rclumpsubscript𝑅clumpR_{\rm clump}italic_R start_POSTSUBSCRIPT roman_clump end_POSTSUBSCRIPT from Liu et al. (2023b) and plot them (orange stars) with ACA fragments (blue pentagons) in Figure 4. The source extraction algorithm of the clumps is the same as we do for ACA fragments, so there is no systematic bias due to the methodology in the following discussion. The mass versus radius (M𝑀Mitalic_MR𝑅Ritalic_R) diagram can be used to study mass concentration at different scales. Sources that shares the same density profile should exhibit scaling relations with the same power law index in the M𝑀Mitalic_MR𝑅Ritalic_R diagram.

We perform a linear regression on the M𝑀Mitalic_MR𝑅Ritalic_R diagram for the QUARKS clumps, resulting in a correlation of MR1.8proportional-to𝑀superscript𝑅1.8M\propto R^{1.8}italic_M ∝ italic_R start_POSTSUPERSCRIPT 1.8 end_POSTSUPERSCRIPT (shown as orange line) with correlation coefficient of 0.91 and 1σ1𝜎1\sigma1 italic_σ data scatter of 0.26. Assuming that all the QUARKS clumps share a similar density profile, then the power-law index corresponds to the expectation of the turbulent-support model proposed by Li (2017). In their model, energy dissipation rate of external turbulence balances with that of internal virialized turbulence, resulting in a mass concentration relationship described by MR1.67proportional-to𝑀superscript𝑅1.67M\propto R^{1.67}italic_M ∝ italic_R start_POSTSUPERSCRIPT 1.67 end_POSTSUPERSCRIPT.

The agreement between observations and theory suggests that the QUARKS massive clumps are currently in a transitional phase. In a turbulence-dominated cloud, a shallower density profile following ρR1proportional-to𝜌superscript𝑅1\rho\propto R^{-1}italic_ρ ∝ italic_R start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT (Butler & Tan 2012) and a steeper mass concentration of M(<r)R2proportional-toannotated𝑀absent𝑟superscript𝑅2M(<r)\propto R^{2}italic_M ( < italic_r ) ∝ italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT are expected. In contrast, in a dense structure dominated by gravity and the system reaches a quasi-stationary stage (Xu et al. 2023a) after relaxation processes, the density profile usually adheres to ρR2proportional-to𝜌superscript𝑅2\rho\propto R^{-2}italic_ρ ∝ italic_R start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT (Li 2018), with the mass concentration described by M(<R)Rproportional-toannotated𝑀absent𝑅𝑅M(<R)\propto Ritalic_M ( < italic_R ) ∝ italic_R.

As shown in blue pentagons of Figure 4, the ACA fragments show a notable deviation from clump (orange color) at scales 0.1absent0.1\leq 0.1≤ 0.1 pc, suggesting a different mass-radius relationship. These fragments have a higher mass for a given radius compared to the turbulent-supported model. A linear regression applied to the ACA fragments reveals a correlation of MR1.1proportional-to𝑀superscript𝑅1.1M\propto R^{1.1}italic_M ∝ italic_R start_POSTSUPERSCRIPT 1.1 end_POSTSUPERSCRIPT, represented by a blue line, suggesting a power-law index close to 1. This scaling aligns with a density profile of ρR2proportional-to𝜌superscript𝑅2\rho\propto R^{-2}italic_ρ ∝ italic_R start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT, indicative of a scale-free gravitational collapse in a self-similar fashion (Li 2018). In Section 5.3.1, self-similarity of ACA fragments will be proposed in another manner.

The transition of mass concentration from several parsecs to tenths parsec is highly consistent with what has been found by Peretto et al. (2023), where star cluster progenitors are reported to be dynamically decoupled from their parent molecular clouds, exhibiting steeper density profiles ρR2proportional-to𝜌superscript𝑅2\rho\propto R^{-2}italic_ρ ∝ italic_R start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT and flat velocity dispersion profiles σR0proportional-to𝜎superscript𝑅0\sigma\propto R^{0}italic_σ ∝ italic_R start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT, clearly departing from Larson’s relations. Similar scale-dependent gas dynamics have been found in several cases of multi-scale studies (Liu et al. 2022; Saha et al. 2022), where gravity-driven gas motion takes over turbulence. In the QUARKS sample, it would be of great interest to investigate the behaviour of M𝑀Mitalic_MR𝑅Ritalic_R down to dense core scale as a follow-up work, similar to what has been done in infrared dark clouds (e.g. Li et al. 2023).

5.1.3 Protocluster ensembles

As shown in high-resolution studies by Xu et al. (2024b) and Liu et al. (2023b), the QUARKS ACA 1.3 mm continuum sources contain protoclusters with a large number of dense cores, which are embedded in the parent ACA sources (see also Zhang et al. 2021). To further demonstrate this ubiquity, we performed a spatial cross-match between the ATOMS 3 mm continuum dense cores by Liu et al. (2021) and the QUARKS ACA 1.3 mm continuum sources.

As a result, we have identified a total of 301 “QUARKS-ATOMS links” (links hereafter), as defined by Eq.11 discussed in Appendix C and listed them in Table 4. The remaining 128 ATOMS dense cores, without any associated ACA sources, are referred to as “field sources”. The presence of field sources can be attributed to the generally higher mass sensitivity of the ATOMS data compared to the QUARKS ACA data. Given a distance of 3 kpc and dust temperature of 20 K, typical ATOMS and QUARKS 7-m continuum sensitivities of 0.2 and 15 mJy give sensitivity limits of 1.5 and 3 Msubscript𝑀direct-productM_{\odot}italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, respectively. Compared to 1.3 mm, the 3 mm continuum emission can be contaminated by free-free emission more easily and diffuse 3 mm emission cannot be seen in the ACA 1.3 mm continuum images.

There are 86 ACA sources (42%) with more than one ATOMS dense cores, and among them, 25 have three or more ATOMS dense cores, indicating the presence of substructures. Consequently, the detected QUARKS ACA sources are likely to be ensembles of protoclusters in nature. However, there are 95 ACA sources associated with single ATOMS dense cores, possibly due to limited resolution. Above all, the analyses at the scale of ACA sources provide a global view of massive protoclusters.

5.2 Correlation Between Clumps and Fragments

As discussed in Section 5.1.2, the ACA 1.3 mm continuum sources trace dense fragments (nH2>104subscript𝑛subscriptH2superscript104n_{\rm H_{2}}>10^{4}italic_n start_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT > 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT cm-3) within massive clumps. Figure 5 presents maximum mass of ACA sources (Msource,maxsubscript𝑀sourcemaxM_{\rm source,max}italic_M start_POSTSUBSCRIPT roman_source , roman_max end_POSTSUBSCRIPT) and the total mass of ACA sources (Msource,totalsubscript𝑀sourcetotalM_{\rm source,total}italic_M start_POSTSUBSCRIPT roman_source , roman_total end_POSTSUBSCRIPT) versus their natal clump mass (Mclumpsubscript𝑀clumpM_{\rm clump}italic_M start_POSTSUBSCRIPT roman_clump end_POSTSUBSCRIPT). These are shown with blue stars and orange triangles, respectively. Comparing the two panels, we find no clear difference between Msource,maxsubscript𝑀sourcemaxM_{\rm source,max}italic_M start_POSTSUBSCRIPT roman_source , roman_max end_POSTSUBSCRIPT and Msource,totalsubscript𝑀sourcetotalM_{\rm source,total}italic_M start_POSTSUBSCRIPT roman_source , roman_total end_POSTSUBSCRIPT in most cases, because the most massive ACA sources are dominated by mass within clumps.

Linear regression is used to correlate Msource,maxsubscript𝑀sourcemaxM_{\rm source,max}italic_M start_POSTSUBSCRIPT roman_source , roman_max end_POSTSUBSCRIPT with Mclumpsubscript𝑀clumpM_{\rm clump}italic_M start_POSTSUBSCRIPT roman_clump end_POSTSUBSCRIPT. The derived scaling relation is log(Msource,max/M)=0.96(logMclump/M)1.33subscript𝑀sourcemaxsubscript𝑀direct-product0.96subscript𝑀clumpsubscript𝑀direct-product1.33\log(M_{\rm source,max}/M_{\odot})=0.96(\log M_{\rm clump}/M_{\odot})-1.33roman_log ( italic_M start_POSTSUBSCRIPT roman_source , roman_max end_POSTSUBSCRIPT / italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT ) = 0.96 ( roman_log italic_M start_POSTSUBSCRIPT roman_clump end_POSTSUBSCRIPT / italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT ) - 1.33 as indicated by blue line, with the correlation coefficient of 0.82 and 1σ𝜎\sigmaitalic_σ data scatter of 0.34 dex. To examine the distance effects on the mass correlation, we further perform linear regression in narrower distance bins (see Appendix D).

Consistent with our results, Lin et al. (2019) find Msource,maxMclump0.96proportional-tosubscript𝑀sourcemaxsuperscriptsubscript𝑀clump0.96M_{\rm source,max}\propto M_{\rm clump}^{0.96}italic_M start_POSTSUBSCRIPT roman_source , roman_max end_POSTSUBSCRIPT ∝ italic_M start_POSTSUBSCRIPT roman_clump end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0.96 end_POSTSUPERSCRIPT in the 350 μ𝜇\muitalic_μm observations of 204 ATLASGAL clumps. Besides, Traficante et al. (2023) also find a scaling relation of Msource,maxMclump1.02proportional-tosubscript𝑀sourcemaxsuperscriptsubscript𝑀clump1.02M_{\rm source,max}\propto M_{\rm clump}^{1.02}italic_M start_POSTSUBSCRIPT roman_source , roman_max end_POSTSUBSCRIPT ∝ italic_M start_POSTSUBSCRIPT roman_clump end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1.02 end_POSTSUPERSCRIPT in an ALMA survey of 13 massive clumps. However, the quasi-linear correlation could be a result of an evolutionary process, as it may not be evident in the early stages of massive clumps. For example, samples of massive starless clumps have shown a significantly small amount of mass stored in fragments (Sanhueza et al. 2019; Svoboda et al. 2019; Morii et al. 2023), consistent with the idea that initial fragmentation in massive clumps are Jeans-like and producing low-mass cores. As a clump evolves, the continuous mass accretion feeds the most massive core (Xu et al. 2023a) and the mass correlation builds up (Xu et al. 2024b). The QUARKS sample predominantly covers mid- and late-stage massive star-forming clumps, characterized by luminosity-to-mass ratio range of 4–460 L/Msubscript𝐿direct-productsubscript𝑀direct-productL_{\odot}/M_{\odot}italic_L start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT / italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT and a median value of 35 L/Msubscript𝐿direct-productsubscript𝑀direct-productL_{\odot}/M_{\odot}italic_L start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT / italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, so the mass correlation is expected.

As indicated by orange line in the right panel of Figure 5, linear regression gives log(Msource,total/M)=0.94(logMclump/M)1.22subscript𝑀sourcetotalsubscript𝑀direct-product0.94subscript𝑀clumpsubscript𝑀direct-product1.22\log(M_{\rm source,total}/M_{\odot})=0.94(\log M_{\rm clump}/M_{\odot})-1.22roman_log ( italic_M start_POSTSUBSCRIPT roman_source , roman_total end_POSTSUBSCRIPT / italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT ) = 0.94 ( roman_log italic_M start_POSTSUBSCRIPT roman_clump end_POSTSUBSCRIPT / italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT ) - 1.22. Intriguingly, the correlation is even tighter with coefficient of 0.85 and 1σ𝜎\sigmaitalic_σ data scatter of 0.30 dex, compared to the that of Msource,maxsubscript𝑀sourcemaxM_{\rm source,max}italic_M start_POSTSUBSCRIPT roman_source , roman_max end_POSTSUBSCRIPTMclumpsubscript𝑀clumpM_{\rm clump}italic_M start_POSTSUBSCRIPT roman_clump end_POSTSUBSCRIPT. The tightening correlation indicates that total ACA source (dense gas) mass could be a physical value that is more directly correlated with clump mass. In other words, the Msource,maxsubscript𝑀sourcemaxM_{\rm source,max}italic_M start_POSTSUBSCRIPT roman_source , roman_max end_POSTSUBSCRIPTMclumpsubscript𝑀clumpM_{\rm clump}italic_M start_POSTSUBSCRIPT roman_clump end_POSTSUBSCRIPT relation could be a combined result of Msource,totalsubscript𝑀sourcetotalM_{\rm source,total}italic_M start_POSTSUBSCRIPT roman_source , roman_total end_POSTSUBSCRIPTMclumpsubscript𝑀clumpM_{\rm clump}italic_M start_POSTSUBSCRIPT roman_clump end_POSTSUBSCRIPT and a mass function which correlates Msource,totalsubscript𝑀sourcetotalM_{\rm source,total}italic_M start_POSTSUBSCRIPT roman_source , roman_total end_POSTSUBSCRIPT with Msource,maxsubscript𝑀sourcemaxM_{\rm source,max}italic_M start_POSTSUBSCRIPT roman_source , roman_max end_POSTSUBSCRIPT (e.g., Bonnell et al. 2004; Weidner et al. 2013).

Refer to caption
Figure 5: Maximum ACA source mass Msource,maxsubscript𝑀sourcemaxM_{\rm source,max}italic_M start_POSTSUBSCRIPT roman_source , roman_max end_POSTSUBSCRIPT and total ACA source mass Msource,totalsubscript𝑀sourcetotalM_{\rm source,total}italic_M start_POSTSUBSCRIPT roman_source , roman_total end_POSTSUBSCRIPT versus QUARKS clump mass Mclumpsubscript𝑀clumpM_{\rm clump}italic_M start_POSTSUBSCRIPT roman_clump end_POSTSUBSCRIPT are shown in the left and right panels, respectively. The dashed lines label the cases where Msource,maxsubscript𝑀sourcemaxM_{\rm source,max}italic_M start_POSTSUBSCRIPT roman_source , roman_max end_POSTSUBSCRIPT/Msource,totalsubscript𝑀sourcetotalM_{\rm source,total}italic_M start_POSTSUBSCRIPT roman_source , roman_total end_POSTSUBSCRIPT equals to 0.1, 1, 10, and 100 percent of Mclumpsubscript𝑀clumpM_{\rm clump}italic_M start_POSTSUBSCRIPT roman_clump end_POSTSUBSCRIPT. Linear regressions are performed to fit the data in logarithmic space. The derived scaling relations are: 1) logMsource,max=0.96logMclump1.33subscript𝑀sourcemax0.96subscript𝑀clump1.33\log M_{\rm source,max}=0.96\log M_{\rm clump}-1.33roman_log italic_M start_POSTSUBSCRIPT roman_source , roman_max end_POSTSUBSCRIPT = 0.96 roman_log italic_M start_POSTSUBSCRIPT roman_clump end_POSTSUBSCRIPT - 1.33 (blue line), with correlation coefficient of 0.82 and 1σ𝜎\sigmaitalic_σ data scatter of 0.34 dex; 2) logMsource,total=0.94logMclump1.22subscript𝑀sourcetotal0.94subscript𝑀clump1.22\log M_{\rm source,total}=0.94\log M_{\rm clump}-1.22roman_log italic_M start_POSTSUBSCRIPT roman_source , roman_total end_POSTSUBSCRIPT = 0.94 roman_log italic_M start_POSTSUBSCRIPT roman_clump end_POSTSUBSCRIPT - 1.22 (orange line), with correlation coefficient of 0.85 and 1σ𝜎\sigmaitalic_σ data scatter of 0.30 dex.

5.3 Dense Gas Fraction and Its Assembly

Star formation takes place in dense molecular gas. Here we take the ACA 1.3 mm continuum sources as “dense gas” relative to the total clump gas, and define the dense gas fraction (DGF),

DGFMdenseMclump,DGFsubscript𝑀densesubscript𝑀clump\mathrm{DGF}\equiv\frac{M_{\rm dense}}{M_{\rm clump}},roman_DGF ≡ divide start_ARG italic_M start_POSTSUBSCRIPT roman_dense end_POSTSUBSCRIPT end_ARG start_ARG italic_M start_POSTSUBSCRIPT roman_clump end_POSTSUBSCRIPT end_ARG , (10)

where Mdense=clumpMsourcesubscript𝑀densesuperscriptabsentclumpsubscript𝑀sourceM_{\rm dense}=\sum\limits^{\rm\in clump}M_{\rm source}italic_M start_POSTSUBSCRIPT roman_dense end_POSTSUBSCRIPT = ∑ start_POSTSUPERSCRIPT ∈ roman_clump end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT roman_source end_POSTSUBSCRIPT is the total ACA source mass within the clump. Urquhart et al. (2018) performed the photometry from near- to far-infrared data of these massive clumps, by which the spectral energy distribution (SED) are fitted. By this method, the clump mass es, Mclumpsubscript𝑀clumpM_{\rm clump}italic_M start_POSTSUBSCRIPT roman_clump end_POSTSUBSCRIPT, were derived and are scaled with the updated distances in this work.

As discussed in Section 5.2, the linearity in Figure 5 indicates a constant DGF within the QUARKS sample, giving a value directly by its intercept of 101.22superscript101.2210^{-1.22}10 start_POSTSUPERSCRIPT - 1.22 end_POSTSUPERSCRIPT, i.e., 6%. In an alternative definition of core formation efficiency (CFE), a highly consistent median value of 6% is also reported in Traficante et al. (2023), evidently consistent with what has been found here.

Although showing invariance with Mclumpsubscript𝑀clumpM_{\rm clump}italic_M start_POSTSUBSCRIPT roman_clump end_POSTSUBSCRIPT, DGF still has a scatter of 1–10%. To further explore the origins of scatter, DGF versus clump radius (Rclumpsubscript𝑅clumpR_{\rm clump}italic_R start_POSTSUBSCRIPT roman_clump end_POSTSUBSCRIPT) and luminosity-to-mass ratio (L/M𝐿𝑀L/Mitalic_L / italic_M) diagrams are explored in Section 5.3.1 and Section 5.3.2, respectively

5.3.1 Self-similarity in protocluster formation

Refer to caption
Figure 6: Dense gas fraction (DGF) versus clump radius (Rclumpsubscript𝑅clumpR_{\rm clump}italic_R start_POSTSUBSCRIPT roman_clump end_POSTSUBSCRIPT) of (a) the QUARKS sample and (b) the ASHES sample. The hexagons indicate the probability distributions of data points. The colored stars show the median values with errorbars in the Rclumpsubscript𝑅clumpR_{\rm clump}italic_R start_POSTSUBSCRIPT roman_clump end_POSTSUBSCRIPT bins.

As shown in the panel (a) of Figure 6, the blue hexagons indicate the probability distributions of data points in the DGF versus clump radius diagram. The median DGF over Rclumpsubscript𝑅clumpR_{\rm clump}italic_R start_POSTSUBSCRIPT roman_clump end_POSTSUBSCRIPT bin, as indicated by the blue stars connected by line, shows no discernible systematic variations with Rclumpsubscript𝑅clumpR_{\rm clump}italic_R start_POSTSUBSCRIPT roman_clump end_POSTSUBSCRIPT. Linear regression shows a weak correlation with a Spearman correlation coefficient of -0.12, indicating dense gas mass remains nearly constant relative to the clump mass across different scales from several tenths of parsec to several parsec. It’s worth noting that there are some clumps with very low DGF (<<<1%), indicating that QUARKS field of views are not large enough to cover all the dense gas in clumps with large sizes. However, the coverage-limit effect can be neglected in our sample because the QUARKS pointings are biased to the dense regions according to the ATOMS survey.

If we consider the ACA sources as what has been observed at various scales, that is, from parsec-scale clumps to sub-parsec-scale cores, then the multi-scale invariant DGF suggests that the gas tends to condense or fragment into dense structure with some constant ratio in a hierarchical system. It implies a self-similar fragmentation or collapsing mode in protocluster formation, as proposed in some case studies (e.g., Wang et al. 2011, 2014). More importantly, Dib (2023) performed delta-variance spectrum analysis of 15 ALMA-IMF cloud structures (Motte et al. 2022) and discovered a self-similar regime less-than-or-similar-to\lesssim 0.03–0.3 pc. Referring to panel (a) of Figure 6, one can find it highly consistent with the sizes of the ACA sources, representing the most compact clumps within the protocluster forming clouds (Dib 2023). Therefore, our result favors the self-similarity of density structure in massive protoclusters in a dependent way.

We also retrieved 39 ACA 1.3 mm continuum images from the “ALMA Survey of 70 μ𝜇\muitalic_μm Dark High-mass Clumps in Early Stages” (ASHES hereafter Sanhueza et al. 2019; Morii et al. 2023). It is worth noting that the ASHES observations adopted a mosaic mode with larger fields of view than the QUARKS. To maintain consistency, we cropped the ASHES continuum images to the same size as the QUARKS. After adopting the same source extraction algorithm and mass calculation, we derive the DGF of 39 ASHES clumps which are shown with gray hexagons in the panel (b) of Figure 6. The gray stars indicate the median values in corresponding parameter bins. Similarly, the ASHES sample also shows an invariance of DGF with size. Therefore, the self-similarity works in both early and late stages of the evolution of massive protoclusters.

5.3.2 Dense gas grows with evolution

Refer to caption
Figure 7: Dense gas fraction (DGF) versus clump luminosity-to-mass ratio (L/M𝐿𝑀L/Mitalic_L / italic_M), indicative of clump evolutionary stage. Blue and gray hexagons show the probability distribution for the QUARKS and ASHES samples (Morii et al. 2023), with orange stars showing the median values in the parameter bins. The orange stars connected by dashed lines show the median DGF values of the combined sample QUARKS + ASHES in the L/M𝐿𝑀L/Mitalic_L / italic_M bins from 0.04 to 400 L/Msubscript𝐿direct-productsubscript𝑀direct-productL_{\odot}/M_{\odot}italic_L start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT / italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT. Linear regression is adopted for the combined sample, and the orange dashed line, log(DGF)=0.13log(L/M)1.59DGF0.13𝐿𝑀1.59\log({\rm DGF})=0.13\log{(L/M)}-1.59roman_log ( roman_DGF ) = 0.13 roman_log ( italic_L / italic_M ) - 1.59, illustrates an increasing trend of DGF with L/M𝐿𝑀L/Mitalic_L / italic_M. This relationship shows a Spearman correlation coefficient of 0.41 and a p-value of 3.2×1083.2superscript1083.2\times 10^{-8}3.2 × 10 start_POSTSUPERSCRIPT - 8 end_POSTSUPERSCRIPT. The pink squares points are retrieved from the SQUALO project (Traficante et al. 2023), and green pentagons are from the ALMA survey of massive cluster progenitors in ATLASGAL (Csengeri et al. 2017).

In Figure 6, an intriguing feature is a systematically one-time larger DGF in the QUARKS than that in the ASHES clumps, and the DGF difference seems to be invariant with the clump size. Therefore, we refer to the evolutionary explanation of the DGF difference between two samples. Theoretically, a high-mass star grows in mass during its formation process, so its luminosity should increase. The ratio of bolometric luminosity to envelope mass L/M𝐿𝑀L/Mitalic_L / italic_M, can thus be used to indicate the evolutionary stage (Sridharan et al. 2002; Elia et al. 2017). In Figure 7, the DGF is plotted against L/M𝐿𝑀L/Mitalic_L / italic_M, where the gray and blue hexagons represent the probability distribution of data points of the ASHES and the QUARKS sample, respectively. The orange stars connected by dashed show the DGF median values of the QUARKS+ASHES combined sample in the L/M𝐿𝑀L/Mitalic_L / italic_M bins from 0.04 to 400 L/Msubscript𝐿direct-productsubscript𝑀direct-productL_{\odot}/M_{\odot}italic_L start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT / italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, spanning four orders of magnitude. To demonstrate the evolutionary trend, a linear regression is performed on the data points from the combined sample, resulting in a correlation of log(DGF)=0.13log(L/M)1.59DGF0.13𝐿𝑀1.59\log({\rm DGF})=0.13\log{(L/M)}-1.59roman_log ( roman_DGF ) = 0.13 roman_log ( italic_L / italic_M ) - 1.59, depicted by the orange solid line. The correlation yields a Spearman correlation coefficient of Rs=0.41subscript𝑅𝑠0.41R_{s}=0.41italic_R start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = 0.41 and an associated p𝑝pitalic_p value of 2.9×1082.9superscript1082.9\times 10^{-8}2.9 × 10 start_POSTSUPERSCRIPT - 8 end_POSTSUPERSCRIPT, indicating an evident increase in the mass fraction of the dense part of the clump as it evolves.

We incorporate the DGF results from an ACA 0.87 mm survey (Csengeri et al. 2017, C17 hereafter) and the ”Star formation in QUiescent And Luminous Objects” (SQUALO) project (Traficante et al. 2023, hereafter T23), contributing 30 green and 13 orange data points in Figure 7. To keep consistency, all the clump masses are retrieved from Urquhart et al. (2018) in the following discussion. While the L/M𝐿𝑀L/Mitalic_L / italic_M of the C17 sample spans a similar range to ours, the DGF is systematically larger. This discrepancy arises because C17 adopted a constant temperature Tdust=25subscript𝑇dust25T_{\rm dust}=25italic_T start_POSTSUBSCRIPT roman_dust end_POSTSUBSCRIPT = 25 K throughout the sample, which is systematically lower than what we used in Eq. 1, resulting in a higher mass. T23 improved this method by categorizing the sample into three evolutionary bins and estimating temperatures of 20, 30, and 40 K, respectively. Despite the limited sample size, the wide range of L/Msubscript𝐿direct-productsubscript𝑀direct-productL_{\odot}/M_{\odot}italic_L start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT / italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT in T23 indicates an increasing trend similar to the results of the QUARKS. As a result, two independent data sets mutually verify a dense mass growth in massive star-forming clumps as they evolve.

The methods in both C17 and T23 give a lower temperature compared to ours. According to Equation 1, a lower temperature leads to a higher mass for a given flux. Therefore, we adopt the above temperature estimation methods to our sample, the resulting DGF will be even larger and the increasing trend will become more significant.

The mass growth of dense cores is widely found during the evolution of massive star-forming clumps (Anderson et al. 2021; Traficante et al. 2023; Liu et al. 2023a; Li et al. 2023; Xu et al. 2024b). Very recently, the ASHES team has found an increase in the mass dynamic range with respect to the protostellar core fraction, serving as a proxy for evolutionary stages (Morii et al. 2024). Our work, encompassing a significantly larger sample with a broad range of evolutionary stages, substantiates the continuous growth of dense mass over time. Overall, recent ALMA studies portray a dynamic scenario wherein dense gas accumulates throughout the evolutionary process.

We note that the evolution-dependent DGF cannot explain the total scatter. The wide range of DGFs among the QUARKS samples can also arise from the dynamic balance between gas depletion/stellar feedback and gas infall. Specifically, gas infall and accretion processes function to concentrate the dense gas, while star formation depletes the dense gas. Stellar feedback mechanisms, such as winds and outflows, play a dual role by releasing gas back to the parent clump and preventing further gas accretion. But for the surrounding embedded dense gas structures, their kinematic properties may be less influenced by feedback from the most evolved stars in clumps (Zhou et al. 2023). At any rate, a systematic investigations into gas kinematics and energetics have the potential to unveil dynamic effects on DGF and elucidate the origin of the scatter in DGF values.

5.4 Limited Fragmentation

ACA sources are fragments from the clump scale. The facts that the mean number of fragments per clump is N¯frag1.5similar-tosubscript¯𝑁frag1.5\bar{N}_{\rm frag}\sim 1.5over¯ start_ARG italic_N end_ARG start_POSTSUBSCRIPT roman_frag end_POSTSUBSCRIPT ∼ 1.5 and that 93 clumps have only one fragment, both suggest limited fragmentation, which is consistent with what has been found in another ACA survey by C17.

C17 discussed the possibility that global collapse at the clump scale could explain the excessive mass of the subclump reservoir (Schneider et al. 2010; Peretto et al. 2013; Xu et al. 2023a). This concept implies that the entire clump is involved in a dynamic process, wherein fragments and low-density gas experience global collapse. Equilibrium may never be reached at subclump scales, which aligns with the limited fragmentation observed. C17 observed that the majority of the clumps are likely not in virial equilibrium, suggesting collapse on the clump scale. Furthermore, continuous mass accretion beyond the clump to the core-scale feed could fuel the formation of the protocluster (Avison et al. 2021; Xu et al. 2023a; Yang et al. 2023; Xu et al. 2023b). In this scenario, an increase in the number of fragments with time and a Jeans-like fragmentation to develop in more evolved stages are expected (Palau et al. 2015), and appear to be in conflict with our observed evolution-independent limited fragmentation. However, the conflict can be reconciled, because the QUARKS ACA observations have limited mass sensitivity of 3greater-than-or-equivalent-toabsent3\gtrsim 3≳ 3Msubscript𝑀direct-productM_{\odot}italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, at a temperature of 20 K and distance of 4 kpc and limited spatial resolution of 0.1greater-than-or-equivalent-toabsent0.1\gtrsim 0.1≳ 0.1 pc. Therefore, those small low-mass fragments cannot be effectively resolved even if they exist. Liu et al. (2023b) show an example of Sgr B2(M) in the QUARKS high-resolution (0.3\arcsecsimilar-toabsent0.3\arcsec\sim 0.3\arcsec∼ 0.3) data. Compared to the only detection in the ACA data in I17441-2822, Liu et al. (2023b) identified more than 30 spatially associated cores with the ACA 1.3 mm source and up to 93 cores in the entire cluster of cores.

Nevertheless, we propose another possibility where ACA fragments are the products of turbulent Jeans fragmentation from the natal clump. Although the thermal Jeans mass in massive clumps is as low as several Msubscript𝑀direct-productM_{\odot}italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, at clump density of n¯cl5×1045similar-tosubscript¯𝑛cl5superscript1045\bar{n}_{\rm cl}\sim 5\times 10^{4-5}over¯ start_ARG italic_n end_ARG start_POSTSUBSCRIPT roman_cl end_POSTSUBSCRIPT ∼ 5 × 10 start_POSTSUPERSCRIPT 4 - 5 end_POSTSUPERSCRIPT cm-3 and kinetic temperature of Tkin=20Ksubscript𝑇kin20𝐾T_{\rm kin}=20\,Kitalic_T start_POSTSUBSCRIPT roman_kin end_POSTSUBSCRIPT = 20 italic_K, turbulent Jeans fragmentation favors more massive fragment, larger Jeans length, and therefore less fragments in a clump.

6 Summary

The QUARKS survey, standing for ‘Querying Underlying mechanisms of massive star formation with ALMA-Resolved gas Kinematics and Structures’, is observing 139 massive gas clumps at ALMA Band 6 (λsimilar-to𝜆absent\lambda\simitalic_λ ∼ 1.3 mm). This paper introduces the Atacama Compact Array (ACA) 7-m data of the QUARKS survey, describing the ACA observations and data reduction. Combining multi-wavelength data, we provide the first edition of QUARKS atlas, offering insights into the multiscale and multiphase interstellar medium (ISM) in high-mass star formation.

Leveraging the QUARKS ACA data, we construct the ACA 1.3 mm continuum source catalog with 207 sources. At least one source and up to five sources are found in one clump. Three source-averaged formaldehyde transition lines p-H2CO (3–2) are fitted using non-LTE radiative transfer model, to obtain the gas kinetic temperature and line width. Based on the geometric and flux measurements of the ACA sources, and assuming that gas temperature equals the dust temperature, physical parameters including mass and surface/volume densities are derived. The thermal and non-thermal dispersion, as well as the Mach number, are also calculated from fitted p-H2CO (3–2) line width.

Statistically speaking, the nature of ACA sources is massive gravity-dominated fragments with at the subclump scale, with supersonic turbulence and possibly embedded star-forming protocluster. A quasi-linear correlation between clump mass and ACA source mass is found, which can be explained by the dynamic coevolution between clump and core in a late stage. The dense gas fraction (DGF) is defined as total ACA source mass over the clump mass, and is found to be about 6%, although with a wide scatter of 1–10%. If we consider the massive clump sample as what has been observed at various scales, then the size-independent DGF indicates that the gas conversion efficiency at each scale level remains constant in a hierarchical system, implying a self-similar fragmentation or collapsing mode in protocluster formation. With the data across four orders of magnitude of luminosity-to-mass ratio, we find a significantly increasing trend of DGF with clump evolution. The fragmentation on the subclump scale is limited and the reasons could be that equilibrium may never be reached at subclump scale and massive clumps are undergoing a global collapse.

Acknowledgements.
We thank the anonymous referee for helpful comments. This work has been supported by the National Science Foundation of China (12033005), the National Key R&D Program of China (No. 2022YFA1603102), the China Manned Space Project (CMS-CSST-2021-A09, CMS-CSST-2021-B06), and the China-Chile Joint Research Fund (CCJRF No. 2211). CCJRF is provided by Chinese Academy of Sciences South America Center for Astronomy (CASSACA) and established by National Astronomical Observatories, Chinese Academy of Sciences (NAOC) and Chilean Astronomy Society (SOCHIAS) to support China-Chile collaborations in astronomy. We acknowledge support from the Tianchi Talent Program of Xinjiang Uygur Autonomous Region. This research was carried out in part at the Jet Propulsion Laboratory, California Institute of Technology, under a contract with the National Aeronautics and Space Administration (80NM0018D0004). PS was partially supported by a Grant-in-Aid for Scientific Research (KAKENHI Number JP22H01271 and JP23H01221) of JSPS. AS, DM, GG, and LB gratefully acknowledge support by ANID Basal project FB210003. MJ acknowledges the support of the Research Council of Finland Grant No. 348342. K.M is supported by a Grants-in-Aid for the the JSPS Fellows (KAKENHI Number 22J21529). Data analysis was in part carried out on the Multi-wavelength Data Analysis System operated by the Astronomy Data Center (ADC), National Astronomical Observatory of Japan. This paper uses the following ALMA data: ADS/JAO.ALMA#2019.1.00685.S and 2021.1.00095.S. ALMA is a partnership of 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 MeerKAT telescope is operated by the South African Radio Astronomy Observatory, which is a facility of the National Research Foundation, an agency of the Department of Science and Innovation. Software. This research uses astropy, a community-developed core Python package for Astronomy (Astropy Collaboration et al. 2013, 2018, 2022). This research has used the program SExtractor, which builds a catalog of objects from an astronomical image (Bertin & Arnouts 1996). This research has used Python-based package pyspeckit to fit spectral lines (Ginsburg & Mirocha 2011; Ginsburg et al. 2022). This research has used RADEX, a computer program for fast non-LTE analysis of interstellar line spectra (van der Tak et al. 2007).

Appendix A Source Extraction

For each field, we initially generated the background and rms map from the original continuum map without applying primary beam correction (unpbcor). Within each unit of boxes, whose size is equivalent to the maximum recoverable scale, we performed iterative clipping of the local background histogram until convergence was achieved at ±3σplus-or-minus3𝜎\pm 3\sigma± 3 italic_σ around its median (Bertin & Arnouts 1996). The background subtracted continuum map, together with the corresponding rms map, was then used as input for SExtractor. Before the program runs, nthresh=4nthresh4\texttt{nthresh}=4nthresh = 4 is set to mask the low SNR (nthresh×\times×local rms) pixels. During the extraction procedure, the deblending parameters, that is, the number of thresholds for the deblending, and the deblending contrast are set deblend_nthresh=512deblend_nthresh512\texttt{deblend\_nthresh}=512deblend_nthresh = 512 and deblend_cont=105deblend_contsuperscript105\texttt{deblend\_cont}=10^{-5}deblend_cont = 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT. To ensure that the extraction focused on genuine sources rather than spurious features or cleaning artifacts, we set the parameter controlling the minimum pixel count for a source, minarea, to match the effective beam size.

Due to the Gaussian-like primary beam response, the real flux of source should be corrected by primary beam correction (pbcor). So, the pb map is interpolated to the barycenter and at the intensity peak of the source, by which the measured integrated flux and peak intensity are divided, respectively.

Appendix B Non-LTE Model Fitting of H2CO

Following the method introduced in Ginsburg et al. (2016), we used RADEX (van der Tak et al. 2007) to create model grids for the p-H2CO molecular lines over 100 densities of n=102.5𝑛superscript102.5n=10^{2.5}italic_n = 10 start_POSTSUPERSCRIPT 2.5 end_POSTSUPERSCRIPT107superscript10710^{7}10 start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT cm-3, 100 H2CO column densities of N𝑁Nitalic_N(H2CO)=1011absentsuperscript1011=10^{11}= 10 start_POSTSUPERSCRIPT 11 end_POSTSUPERSCRIPT1015superscript101510^{15}10 start_POSTSUPERSCRIPT 15 end_POSTSUPERSCRIPT cm-2, and 50 kinetic temperatures of Tkin=10subscript𝑇kin10T_{\rm kin}=10italic_T start_POSTSUBSCRIPT roman_kin end_POSTSUBSCRIPT = 10350350350350 K, with a fixed assumed line gradient of 5 km s-1 pc-1. In the grid modeling, the collision rates were taken from Wiesenfeld & Faure (2013) and calculated for temperatures in the range from 10 to 300 K including energy levels up to about 200 cm-1 for collisions with H2. Based on the preconstructed non-LTE model grids, the line fitting is performed. The five parameters are kinetic temperature (Tkinsubscript𝑇kinT_{\rm kin}italic_T start_POSTSUBSCRIPT roman_kin end_POSTSUBSCRIPT), formaldehyde column density (logN(H2CO)𝑁subscriptH2CO\log N(\rm H_{2}CO)roman_log italic_N ( roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT roman_CO )), hydrogen molecule volume density (logn(H2)𝑛subscriptH2\log n(\rm H_{2})roman_log italic_n ( roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT )), centroid velocity (vlsrsubscript𝑣lsrv_{\rm lsr}italic_v start_POSTSUBSCRIPT roman_lsr end_POSTSUBSCRIPT) and line width (FWHM). For all the sources, only one velocity component is considered.

Refer to caption
Figure 8: As an example, the spectral line data and the fitting model of I13291-6229_ACA2 are shown in red and black color, respectively. The fitted parameters kinetic temperature Tkinsubscript𝑇kinT_{\rm kin}italic_T start_POSTSUBSCRIPT roman_kin end_POSTSUBSCRIPT (K), formaldehyde column density logN(H2CO)𝑁subscriptH2CO\log N(\rm H_{2}CO)roman_log italic_N ( roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT roman_CO ) (cm-2), hydrogen molecule volume density logn(H2)𝑛subscriptH2\log n(\rm H_{2})roman_log italic_n ( roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) (cm-3), centroid velocity vlsrsubscript𝑣lsrv_{\rm lsr}italic_v start_POSTSUBSCRIPT roman_lsr end_POSTSUBSCRIPT (km s-1) and FWHM line width (km s-1) are shown on the top right. The unfitted line between 218.4 GHz and 218.5 GHz is CH3OH JK=42,231,2subscript𝐽𝐾subscript422subscript312J_{K}=4_{2,2}-3_{1,2}italic_J start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT = 4 start_POSTSUBSCRIPT 2 , 2 end_POSTSUBSCRIPT - 3 start_POSTSUBSCRIPT 1 , 2 end_POSTSUBSCRIPT.

An example (I13291-6229_ACA2) of spectral line data and the fitting model is shown in Figure 8, where Tkinsubscript𝑇kinT_{\rm kin}italic_T start_POSTSUBSCRIPT roman_kin end_POSTSUBSCRIPT is estimated to be 53(±11plus-or-minus11\pm 11± 11) K. If the fitting fails (for seven spectra), then the Tkinsubscript𝑇kinT_{\rm kin}italic_T start_POSTSUBSCRIPT roman_kin end_POSTSUBSCRIPT is set to be equal to clump-averaged dust temperature Tdust,clumpsubscript𝑇dustclumpT_{\rm dust,clump}italic_T start_POSTSUBSCRIPT roman_dust , roman_clump end_POSTSUBSCRIPT which is retrieved from Urquhart et al. (2018). The kinetic temperatures for 207 ACA sources are then listed in column (3) of Table 3.

B.1 Validation

The volume density of the collisional partner, which is molecular hydrogen (H2) in our case, serves as one of the non-LTE model parameters. Assuming a dust emission model and a good mix of dust and gas, it is possible to determine the source-averaged volume density of H2. This value is listed in column (8) of Table 3 independently.

As depicted in panel (f) of Figure 3, the density of ACA sources predominantly falls within the range of 104superscript10410^{4}10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT to 106superscript10610^{6}10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT cm-3, which is well-suited for H2CO triplet model fitting. When the density surpasses 106superscript10610^{6}10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT cm-3, the line ratio becomes less sensitive to kinetic temperature, primarily due to the effects of radiative trapping, as shown in Figure 6 of Ao et al. (2013). As we revisit and scrutinize the input parameters, we ensure the validity of our non-LTE model fitting for most of the QUARKS ACA sources.

In Figure 9, we compare the volume density derived from RADEX line modeling nH2,RADEXsubscript𝑛subscriptH2RADEXn_{\rm H_{2},RADEX}italic_n start_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , roman_RADEX end_POSTSUBSCRIPT and that derived from dust emission nH2,dustsubscript𝑛subscriptH2dustn_{\rm H_{2},dust}italic_n start_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , roman_dust end_POSTSUBSCRIPT (Section 4.3). The data points follow a bulk increasing trend although with a large dispersion, further justifying the self-consistency of the temperature estimation. Besides, we note that nH2,RADEXsubscript𝑛subscriptH2RADEXn_{\rm H_{2},RADEX}italic_n start_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , roman_RADEX end_POSTSUBSCRIPT is systematically higher than nH2,dustsubscript𝑛subscriptH2dustn_{\rm H_{2},dust}italic_n start_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , roman_dust end_POSTSUBSCRIPT, as depicted by black dashed lines. This can be explained by the difference in spatial distribution or the sizes that the H2CO/dust trace. If the size that dust traces is higher than that H2CO traces, then nH2,dustsubscript𝑛subscriptH2dustn_{\rm H_{2},dust}italic_n start_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , roman_dust end_POSTSUBSCRIPT should be naturally lower than nH2,RADEXsubscript𝑛subscriptH2RADEXn_{\rm H_{2},RADEX}italic_n start_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , roman_RADEX end_POSTSUBSCRIPT. Therefore, QUARKS high-resolution data should be essential to resolving and understanding internal density structure.

Refer to caption
Figure 9: Volume density nH2subscript𝑛subscriptH2n_{\rm H_{2}}italic_n start_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT derived from RADEX line modeling versus that derived from dust emission in a sample of ACA sources with RADEX modeling. The black dashed lines mark nH2,dust=0.1/1/10×nH2,RADEXsubscript𝑛subscriptH2dust0.1110subscript𝑛subscriptH2RADEXn_{\rm H_{2},dust}=0.1/1/10\times n_{\rm H_{2},RADEX}italic_n start_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , roman_dust end_POSTSUBSCRIPT = 0.1 / 1 / 10 × italic_n start_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , roman_RADEX end_POSTSUBSCRIPT.

B.2 Caveats

The temperature estimation using the H2CO triplet assumes that the H2CO emission mainly traces dense gas, which is not the case in a sample of 12 infrared dark clouds (Izumi et al. 2023). These authors argued that H2CO emission is mainly sensitive to low-velocity outflow components rather than to quiescent gas expected in the early phases of star formation. So one must be keep in mind the fidelity of the temperature estimated from H2CO triplets depends on how well H2CO emission traces the gas in dense cores.

The non-LTE line modeling can be influenced by the line profile, including the wings of the line and self-absorption. For example, line wings can make least-square method fall into a false local minimum. There are six cases of severe self-absorption, for which we only assign them Tdust,clumpsubscript𝑇dustclumpT_{\rm dust,clump}italic_T start_POSTSUBSCRIPT roman_dust , roman_clump end_POSTSUBSCRIPT. If more than one velocity component is associated with an ACA source, the line width will be overestimated.

Appendix C QUARKS-ATOMS Link

Elliptical mask is defined by the geometrical measurement of the ACA 1.3 mm continuum source,

[(xxQ)cosϕ+(yyQ)sinϕ]2A2+[(xxQ)sinϕ(yyQ)cosϕ]2B21,superscriptdelimited-[]𝑥subscript𝑥𝑄italic-ϕ𝑦subscript𝑦𝑄italic-ϕ2superscript𝐴2superscriptdelimited-[]𝑥subscript𝑥𝑄italic-ϕ𝑦subscript𝑦𝑄italic-ϕ2superscript𝐵21\begin{split}&\frac{\left[(x-x_{Q})\cos\phi+(y-y_{Q})\sin\phi\right]^{2}}{A^{2% }}\\ &+\frac{\left[(x-x_{Q})\sin\phi-(y-y_{Q})\cos\phi\right]^{2}}{B^{2}}\leq 1,% \end{split}start_ROW start_CELL end_CELL start_CELL divide start_ARG [ ( italic_x - italic_x start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT ) roman_cos italic_ϕ + ( italic_y - italic_y start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT ) roman_sin italic_ϕ ] start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_A start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL + divide start_ARG [ ( italic_x - italic_x start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT ) roman_sin italic_ϕ - ( italic_y - italic_y start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT ) roman_cos italic_ϕ ] start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ≤ 1 , end_CELL end_ROW (11)

where x,y𝑥𝑦{x,y}italic_x , italic_y and xQ,yQsubscript𝑥𝑄subscript𝑦𝑄{x_{Q},y_{Q}}italic_x start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT , italic_y start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT are respectively the coordinates of the ATOMS and QUARKS sources, A𝐴Aitalic_A, B𝐵Bitalic_B, and ϕitalic-ϕ\phiitalic_ϕ are the major axis, minor axis, and the position angle of the QUARKS source. All parameters can be found in Table 2 and ATOMS Table 4. A QUARKS-ATOMS source link (link hereafter) is established when Eq. 11 is satisfied.

We list the link, as introduced in Section 5.1, between ATOMS 3 mm dense cores and QUARKS 1.3 mm sources in Table 4. The field name and the ATOMS dense core ID are listed in columns (1)–(2). The Galactic name of dense core, inherited from Liu et al. (2021), is listed in column (3). The ICRS coordinates of the barycenter are listed in columns (4)–(5). The associated QUARKS source ID is listed in column (6). If no associated QUARKS source, then the ATOMS sources are referred as “field sources”, which are marked by “0”. The angular size (Ldsubscript𝐿𝑑L_{d}italic_L start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT) and position angle (LPAsubscript𝐿PAL_{\rm PA}italic_L start_POSTSUBSCRIPT roman_PA end_POSTSUBSCRIPT) of the link are listed in columns (7)–(8). The link fidelity, defined as how close the ATOMS source is to the QUARKS source, with values from 0 (unreliable) to 1 (reliable), is listed in column (9).

Table 4: ATOMS 3 mm Dense Cores Linked to QUARKS 1.3 mm ACA Sources .
Field ATOMS ID Galactic Name Equatorial Coordinates QUARKS ID Ldsubscript𝐿𝑑L_{d}italic_L start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT LPAsubscript𝐿PAL_{\rm PA}italic_L start_POSTSUBSCRIPT roman_PA end_POSTSUBSCRIPT Fidelity
RA (ICRS) DEC (ICRS) (\arcsec) ()
(1) (2) (3) (4) (5) (6) (7) (8) (9)
I08303-4303 1 G261.6444-02.0876 08:32:09.0 -43:13:42.9 1 4.5 41.6 0.69
I08303-4303 2 G261.6444-02.0890 08:32:08.6 -43:13:45.8 1 1.2 108.5 0.99
I08303-4303 3 G261.6446-02.0899 08:32:08.4 -43:13:48.3 1 3.9 57.0 0.8
I08448-4343 1 G263.7745-00.4266 08:46:35.0 -43:54:23.8 1 1.4 28.5 0.93
I08448-4343 2 G263.7756-00.4281 08:46:34.9 -43:54:30.3 2 3.5 23.7 0.57
I08448-4343 3 G263.7756-00.4291 08:46:34.6 -43:54:32.6 2 1.7 123.7 0.94
I08448-4343 4 G263.7743-00.4317 08:46:33.7 -43:54:34.8 3 3.5 59.9 0.77
I08448-4343 5 G263.7737-00.4326 08:46:33.3 -43:54:35.1 3 1.7 150.5 0.91
I08448-4343 6 G263.7712-00.4363 08:46:31.8 -43:54:36.4 4 7.6 95.0 0.59
I08448-4343 7 G263.7723-00.4350 08:46:32.4 -43:54:36.6 4 1.6 108.5 0.97
I08448-4343 8 G263.7700-00.4379 08:46:31.1 -43:54:36.7 0
I08448-4343 9 G263.7744-00.4328 08:46:33.4 -43:54:37.5 3 1.0 168.4 0.96
I08448-4343 10 G263.7766-00.4309 08:46:34.3 -43:54:39.4 0
I08448-4343 11 G263.7724-00.4366 08:46:32.0 -43:54:40.5 4 6.7 59.6 0.19
I08448-4343 12 G263.7729-00.4364 08:46:32.1 -43:54:41.4 0
I08448-4343 13 G263.7776-00.4332 08:46:34.0 -43:54:47.4 0

The field name and the ATOMS dense core ID are listed in columns (1)–(2). The Galactic name of dense core is listed in column (3). The ICRS coordinates of the barycenter are listed in columns (4)–(5). The associated QUARKS source ID is listed in column (6). If there is no associated QUARKS source, then “0”. The angular size (Ldsubscript𝐿𝑑L_{d}italic_L start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT) and position angle (LPAsubscript𝐿PAL_{\rm PA}italic_L start_POSTSUBSCRIPT roman_PA end_POSTSUBSCRIPT) of the link are listed in columns (7)–(8). The fidelity of the link is listed in column (9). The table is available in its entirety in machine-readable form.

Note that there are some ACA sources that lack associated ATOMS 3 mm sources, which can be categorized into two primary scenarios. In one scenario, the ATOMS dense core catalog prioritizes the high fidelity of genuinely dense cores, but this can lead to a higher false negative rate. Consequently, some QUARKS ACA sources with non-spherical morphologies may be missed in the ATOMS dense core catalog. An example of this is seen in sources I16132-5039_ACA1 and I16132-5039_ACA2, which are elongated sources without any associated ATOMS dense cores. In the other scenario, QUARKS data, particularly in certain fields, exhibit greater sensitivity than ATOMS. This higher sensitivity enables the detection of fainter structures. For example, I17269-3312_ACA3 and I17269-3312_ACA4, two relatively faint sources in the I17269-3312 field, are marginally seen in the ATOMS data and are not included in the catalog.

Appendix D Distance Effects on Mass Correlation

The calculations of the clump and the ACA source masses depend on distance by Md2proportional-to𝑀superscript𝑑2M\propto d^{2}italic_M ∝ italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. Therefore, it is crucial to examine the impact of distance on the mass correlation discussed in Section 5.2. To explore this, we categorize the QUARKS clumps into three groups based on their distances: d3𝑑3d\leq 3italic_d ≤ 3 kpc (Near), 3<d63𝑑63<d\leq 63 < italic_d ≤ 6 kpc (Mid), and d>6𝑑6d>6italic_d > 6 kpc (Far). The selection of distance bins is solely based on achieving a similar sample size in each group.

We perform linear regression on the three groups individually and obtain similar quasi-linear correlations as observed in the total sample. The correlation coefficients for the three fittings are substantial, ranging from 0.72 to 0.74, and the 1σ1𝜎1\sigma1 italic_σ scatters fall between 0.25 and 0.35. Both metrics indicate a significant correlation between Mclumpsubscript𝑀clumpM_{\rm clump}italic_M start_POSTSUBSCRIPT roman_clump end_POSTSUBSCRIPT and Msource,totalsubscript𝑀sourcetotalM_{\rm source,total}italic_M start_POSTSUBSCRIPT roman_source , roman_total end_POSTSUBSCRIPT. Additionally, we observe wide mass range for narrow color (i.e., distance) range, suggesting that even narrower distance bins would still result in a relatively strong mass correlation. In summary, we assert a weak distance effect on the mass correlation in the QUARKS sample.

Refer to caption
Figure 10: The correlation between the QUARKS clump mass Mclumpsubscript𝑀clumpM_{\rm clump}italic_M start_POSTSUBSCRIPT roman_clump end_POSTSUBSCRIPT and the total ACA source mass Msource,totalsubscript𝑀sourcetotalM_{\rm source,total}italic_M start_POSTSUBSCRIPT roman_source , roman_total end_POSTSUBSCRIPT in different distance bins of d3𝑑3d\leq 3italic_d ≤ 3 kpc (left), 3<d63𝑑63<d\leq 63 < italic_d ≤ 6 kpc (middle), and d>6𝑑6d>6italic_d > 6 kpc (right). The dashed lines label the case where Msource,maxsubscript𝑀sourcemaxM_{\rm source,max}italic_M start_POSTSUBSCRIPT roman_source , roman_max end_POSTSUBSCRIPT/Msource,totalsubscript𝑀sourcetotalM_{\rm source,total}italic_M start_POSTSUBSCRIPT roman_source , roman_total end_POSTSUBSCRIPT equals to 0.1, 1, 10, and 100 percent of Mclumpsubscript𝑀clumpM_{\rm clump}italic_M start_POSTSUBSCRIPT roman_clump end_POSTSUBSCRIPT. Linear regression is performed to fit the data in logarithmic space, and the fitting result is shown with a blue solid line. The clump distances are coded in the colors of the stars.

References

  • Anderson et al. (2021) Anderson, M., Peretto, N., Ragan, S. E., et al. 2021, MNRAS, 508, 2964
  • Ao et al. (2013) Ao, Y., Henkel, C., Menten, K. M., et al. 2013, A&A, 550, A135
  • Astropy Collaboration et al. (2013) Astropy Collaboration, Robitaille, T. P., Tollerud, E. J., et al. 2013, A&A, 558, A33
  • Astropy Collaboration et al. (2018) Astropy Collaboration, Price-Whelan, A. M., Sipőcz, B. M., et al. 2018, AJ, 156, 123
  • Astropy Collaboration et al. (2022) Astropy Collaboration, Price-Whelan, A. M., Lim, P. L., et al. 2022, ApJ, 935, 167
  • Avison et al. (2015) Avison, A., Peretto, N., Fuller, G. A., et al. 2015, A&A, 577, A30
  • Avison et al. (2021) Avison, A., Fuller, G. A., Peretto, N., et al. 2021, A&A, 645, A142
  • Barnes et al. (2021) Barnes, A. T., Henshaw, J. D., Fontani, F., et al. 2021, MNRAS, 503, 4601
  • Bertin & Arnouts (1996) Bertin, E., & Arnouts, S. 1996, A&AS, 117, 393
  • Bonnell et al. (2004) Bonnell, I. A., Vine, S. G., & Bate, M. R. 2004, MNRAS, 349, 735
  • Butler & Tan (2012) Butler, M. J., & Tan, J. C. 2012, ApJ, 754, 5
  • CASA Team et al. (2022) CASA Team, Bean, B., Bhatnagar, S., et al. 2022, PASP, 134, 114501
  • Chen et al. (2019) Chen, H.-R. V., Zhang, Q., Wright, M. C. H., et al. 2019, ApJ, 875, 24
  • Contreras et al. (2013) Contreras, Y., Schuller, F., Urquhart, J. S., et al. 2013, A&A, 549, A45
  • Contreras et al. (2018) Contreras, Y., Sanhueza, P., Jackson, J. M., et al. 2018, ApJ, 861, 14
  • Cornwell (2008) Cornwell, T. J. 2008, IEEE Journal of Selected Topics in Signal Processing, 2, 793
  • Csengeri et al. (2017) Csengeri, T., Bontemps, S., Wyrowski, F., et al. 2017, A&A, 600, L10
  • Dib (2023) Dib, S. 2023, MNRAS, 524, 1625
  • Elia et al. (2017) Elia, D., Molinari, S., Schisano, E., et al. 2017, MNRAS, 471, 100
  • Evans et al. (2022) Evans, N. J., Kim, J.-G., & Ostriker, E. C. 2022, ApJ, 929, L18
  • Federrath & Klessen (2012) Federrath, C., & Klessen, R. S. 2012, ApJ, 761, 156
  • Ginsburg & Mirocha (2011) Ginsburg, A., & Mirocha, J. 2011, PySpecKit: Python Spectroscopic Toolkit, Astrophysics Source Code Library, record ascl:1109.001
  • Ginsburg et al. (2022) Ginsburg, A., Sokolov, V., de Val-Borro, M., et al. 2022, AJ, 163, 291
  • Ginsburg et al. (2016) Ginsburg, A., Henkel, C., Ao, Y., et al. 2016, A&A, 586, A50
  • Goedhart et al. (2023) Goedhart, S., Cotton, W. D., Camilo, F., et al. 2023, arXiv e-prints, arXiv:2312.07275
  • Goldsmith (2001) Goldsmith, P. F. 2001, ApJ, 557, 736
  • Izumi et al. (2023) Izumi, N., Sanhueza, P., Koch, P. M., et al. 2023, arXiv e-prints, arXiv:2312.03935
  • Kahane et al. (1984) Kahane, C., Frerking, M. A., Langer, W. D., Encrenas, P., & Lucas, R. 1984, A&A, 137, 211
  • Kauffmann et al. (2008) Kauffmann, J., Bertoldi, F., Bourke, T. L., Evans, N. J., I., & Lee, C. W. 2008, A&A, 487, 993
  • Krumholz & McKee (2008) Krumholz, M. R., & McKee, C. F. 2008, Nature, 451, 1082
  • Larson (2003) Larson, R. B. 2003, in Astronomical Society of the Pacific Conference Series, Vol. 287, Galactic Star Formation Across the Stellar Mass Spectrum, ed. J. M. De Buizer & N. S. van der Bliek, 65
  • Li (2017) Li, G.-X. 2017, MNRAS, 465, 667
  • Li (2018) Li, G.-X. 2018, MNRAS, 477, 4951
  • Li et al. (2023) Li, S., Sanhueza, P., Zhang, Q., et al. 2023, ApJ, 949, 109
  • Lin (2021) Lin, Y. 2021, The structure of massive star-forming clumps, PhD thesis, Rheinische Friedrich-Wilhelms-Universität Bonn
  • Lin et al. (2019) Lin, Y., Csengeri, T., Wyrowski, F., et al. 2019, A&A, 631, A72
  • Liu et al. (2021) Liu, H.-L., Liu, T., Evans, Neal J., I., et al. 2021, MNRAS, 505, 2801
  • Liu et al. (2022) Liu, H.-L., Tej, A., Liu, T., et al. 2022, MNRAS, 511, 4480
  • Liu et al. (2023a) Liu, H.-L., Tej, A., Liu, T., et al. 2023a, MNRAS, 522, 3719
  • Liu et al. (2016) Liu, T., Zhang, Q., Kim, K.-T., et al. 2016, ApJ, 824, 31
  • Liu et al. (2020) Liu, T., Evans, N. J., Kim, K.-T., et al. 2020, MNRAS, 496, 2790
  • Liu et al. (2023b) Liu, X., Liu, T., Zhu, L., et al. 2023b, arXiv e-prints, arXiv:2311.08651
  • Lu et al. (2018) Lu, X., Zhang, Q., Liu, H. B., et al. 2018, ApJ, 855, 9
  • Mainzer et al. (2011) Mainzer, A., Bauer, J., Grav, T., et al. 2011, ApJ, 731, 53
  • Mangum & Wootten (1993) Mangum, J. G., & Wootten, A. 1993, ApJS, 89, 123
  • Mathis et al. (1977) Mathis, J. S., Rumpl, W., & Nordsieck, K. H. 1977, ApJ, 217, 425
  • Matzner & McKee (2000) Matzner, C. D., & McKee, C. F. 2000, ApJ, 545, 364
  • McKee & Tan (2003) McKee, C. F., & Tan, J. C. 2003, ApJ, 585, 850
  • 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
  • Morii et al. (2023) Morii, K., Sanhueza, P., Nakamura, F., et al. 2023, ApJ, 950, 148
  • Morii et al. (2024) Morii, K., Sanhueza, P., Zhang, Q., et al. 2024, arXiv e-prints, arXiv:2403.07058
  • Motte et al. (2018) Motte, F., Bontemps, S., & Louvet, F. 2018, ARA&A, 56, 41
  • Motte et al. (2022) Motte, F., Bontemps, S., Csengeri, T., et al. 2022, A&A, 662, A8
  • Müller et al. (2001) Müller, H. S. P., Thorwirth, S., Roth, D. A., & Winnewisser, G. 2001, A&A, 370, L49
  • Neupane et al. (2020) Neupane, S., Garay, G., Contreras, Y., Guzmán, A. E., & Rodríguez, L. F. 2020, ApJ, 890, 76
  • Olguin et al. (2022) Olguin, F. A., Sanhueza, P., Ginsburg, A., et al. 2022, ApJ, 929, 68
  • Ossenkopf & Henning (1994) Ossenkopf, V., & Henning, T. 1994, A&A, 291, 943
  • Padmanabh et al. (2023) Padmanabh, P. V., Barr, E. D., Sridhar, S. S., et al. 2023, arXiv:2303.09231
  • Palau et al. (2014) Palau, A., Estalella, R., Girart, J. M., et al. 2014, ApJ, 785, 42
  • Palau et al. (2015) Palau, A., Ballesteros-Paredes, J., Vázquez-Semadeni, E., et al. 2015, MNRAS, 453, 3785
  • Palau et al. (2021) Palau, A., Zhang, Q., Girart, J. M., et al. 2021, ApJ, 912, 159
  • Peretto et al. (2023) Peretto, N., Rigby, A. J., Louvet, F., et al. 2023, MNRAS, 525, 2935
  • Peretto et al. (2013) Peretto, N., Fuller, G. A., Duarte-Cabral, A., et al. 2013, A&A, 555, A112
  • Peretto et al. (2020) Peretto, N., Rigby, A., André, P., et al. 2020, MNRAS, 496, 3482
  • Pineda et al. (2023) Pineda, J. E., Arzoumanian, D., Andre, P., et al. 2023, in Astronomical Society of the Pacific Conference Series, Vol. 534, Protostars and Planets VII, ed. S. Inutsuka, Y. Aikawa, T. Muto, K. Tomida, & M. Tamura, 233
  • Qin et al. (2022) Qin, S.-L., Liu, T., Liu, X., et al. 2022, MNRAS, 511, 3463
  • Redaelli et al. (2022) Redaelli, E., Bovino, S., Sanhueza, P., et al. 2022, ApJ, 936, 169
  • Reid et al. (2019) Reid, M. J., Menten, K. M., Brunthaler, A., et al. 2019, ApJ, 885, 131
  • Rigby et al. (2021) Rigby, A. J., Peretto, N., Adam, R., et al. 2021, MNRAS, 502, 4576
  • Rosolowsky et al. (2010) Rosolowsky, E., Dunham, M. K., Ginsburg, A., et al. 2010, ApJS, 188, 123
  • Saha et al. (2022) Saha, A., Tej, A., Liu, H.-L., et al. 2022, MNRAS, 516, 1983
  • Sanhueza et al. (2017) Sanhueza, P., Jackson, J. M., Zhang, Q., et al. 2017, ApJ, 841, 97
  • Sanhueza et al. (2019) Sanhueza, P., Contreras, Y., Wu, B., et al. 2019, ApJ, 886, 102
  • Sanhueza et al. (2021) Sanhueza, P., Girart, J. M., Padovani, M., et al. 2021, ApJ, 915, L10
  • Schneider et al. (2010) Schneider, N., Csengeri, T., Bontemps, S., et al. 2010, A&A, 520, A49
  • Schuller et al. (2009) Schuller, F., Menten, K. M., Contreras, Y., et al. 2009, A&A, 504, 415
  • Sridharan et al. (2002) Sridharan, T. K., Beuther, H., Schilke, P., Menten, K. M., & Wyrowski, F. 2002, ApJ, 566, 931
  • Svoboda et al. (2019) Svoboda, B. E., Shirley, Y. L., Traficante, A., et al. 2019, ApJ, 886, 36
  • 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
  • Tang et al. (2017) Tang, X. D., Henkel, C., Chen, C. H. R., et al. 2017, A&A, 600, A16
  • Tang et al. (2021) Tang, X. D., Henkel, C., Menten, K. M., et al. 2021, A&A, 655, A12
  • Traficante et al. (2011) Traficante, A., Calzoletti, L., Veneziani, M., et al. 2011, MNRAS, 416, 2932
  • Traficante et al. (2023) Traficante, A., Jones, B. M., Avison, A., et al. 2023, MNRAS, 520, 2306
  • Urquhart et al. (2014) Urquhart, J. S., Moore, T. J. T., Csengeri, T., et al. 2014, MNRAS, 443, 1555
  • Urquhart et al. (2018) Urquhart, J. S., König, C., Giannetti, A., et al. 2018, MNRAS, 473, 1059
  • van der Tak et al. (2007) van der Tak, F. F. S., Black, J. H., Schöier, F. L., Jansen, D. J., & van Dishoeck, E. F. 2007, A&A, 468, 627
  • Wang (2015) Wang, K. 2015, The Earliest Stages of Massive Clustered Star Formation: Fragmentation of Infrared Dark Clouds
  • Wang et al. (2011) Wang, K., Zhang, Q., Wu, Y., & Zhang, H. 2011, ApJ, 735, 64
  • Wang et al. (2014) Wang, K., Zhang, Q., Testi, L., et al. 2014, MNRAS, 439, 3275
  • Wang et al. (2010) Wang, P., Li, Z.-Y., Abel, T., & Nakamura, F. 2010, ApJ, 709, 27
  • Weidner et al. (2013) Weidner, C., Kroupa, P., & Pflamm-Altenburg, J. 2013, MNRAS, 434, 84
  • Wiesenfeld & Faure (2013) Wiesenfeld, L., & Faure, A. 2013, MNRAS, 432, 2573
  • Wright et al. (2010) Wright, E. L., Eisenhardt, P. R. M., Mainzer, A. K., et al. 2010, AJ, 140, 1868
  • Wu & Evans (2003) Wu, J., & Evans, Neal J., I. 2003, ApJ, 592, L79
  • Wu et al. (2007) Wu, Y., Henkel, C., Xue, R., Guan, X., & Miller, M. 2007, ApJ, 669, L37
  • Xu et al. (2023a) Xu, F.-W., Wang, K., Liu, T., et al. 2023a, MNRAS, 520, 3259
  • Xu et al. (2023b) Xu, F., Wang, K., He, Y., et al. 2023b, ApJS, 269, 38
  • Xu et al. (2021) Xu, F., Wu, Y., Liu, T., et al. 2021, ApJ, 920, 103
  • Xu et al. (2024a) Xu, F., Wang, K., Liu, T., et al. 2024a, ApJ, 963, L9
  • Xu et al. (2024b) Xu, F., Wang, K., Liu, T., et al. 2024b, ApJS, 270, 9
  • Yang et al. (2023) Yang, D., Liu, H.-L., Tej, A., et al. 2023, ApJ, 953, 40
  • Yuan et al. (2018) Yuan, J., Li, J.-Z., Wu, Y., et al. 2018, ApJ, 852, 12
  • Zhang et al. (2009) Zhang, Q., Wang, Y., Pillai, T., & Rathborne, J. 2009, ApJ, 696, 268
  • Zhang et al. (2021) Zhang, S., Zavagno, A., López-Sepulcre, A., et al. 2021, A&A, 646, A25
  • Zhou et al. (2022) Zhou, J.-W., Liu, T., Evans, N. J., et al. 2022, MNRAS, 514, 6038
  • Zhou et al. (2023) Zhou, J. W., Dib, S., Wyrowski, F., et al. 2023, arXiv e-prints, arXiv:2312.01497
  • Zinnecker & Yorke (2007) Zinnecker, H., & Yorke, H. W. 2007, ARA&A, 45, 481