Gas Kinematics of the Massive Protocluster G286.21+0.17 Revealed by ALMA

, , , , and

Published 2020 May 8 © 2020. The American Astronomical Society. All rights reserved.
, , Citation Yu Cheng et al 2020 ApJ 894 87 DOI 10.3847/1538-4357/ab879f

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

0004-637X/894/2/87

Abstract

We study the gas kinematics and dynamics of the massive protocluster G286.21+0.17 with the Atacama Large Millimeter/submillimeter Array using spectral lines of C18O(2–1), ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$(3–2), ${\mathrm{DCO}}^{+}$(3–2), and $\mathrm{DCN}$(3–2). On the parsec clump scale, C18O emission appears highly filamentary around the systemic velocity, ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$ and ${\mathrm{DCO}}^{+}$ are more closely associated with the dust continuum, and $\mathrm{DCN}$ is strongly concentrated toward the protocluster center, where no or only weak detection is seen for ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$ and ${\mathrm{DCO}}^{+}$, possibly due to this region being at a relatively evolved evolutionary stage. Spectra of 76 continuum-defined dense cores, typically a few 1000 au in size, are analyzed to measure their centroid velocities and internal velocity dispersions. There are no statistically significant velocity offsets of the cores among the different dense gas tracers. Furthermore, the majority (71%) of the dense cores have subthermal velocity offsets with respect to their surrounding, lower-density C18O-emitting gas. Within the uncertainties, the dense cores in G286 show internal kinematics that are consistent with being in virial equilibrium. On clump scales, the core-to-core velocity dispersion is also similar to that required for virial equilibrium in the protocluster potential. However, the distribution in velocity of the cores is largely composed of two spatially distinct groups, which indicates that the dense molecular gas has not yet relaxed to virial equilibrium, perhaps due to there being recent/continuous infall into the system.

Export citation and abstract BibTeX RIS

1. Introduction

While it is generally agreed that most stars form in clusters and/or associations rather than in isolation (e.g., Lada & Lada 2003; Gutermuth et al. 2009; Bressert et al. 2010), there is no consensus for how this comes about. Several fundamental questions about star cluster formation are still debated. For example, is the process initiated by internal processes within a giant molecular cloud (GMC), such as decay of support by supersonic turbulence or magnetic fields, or external processes, such as triggering by cloud–cloud collisions or feedback-induced shock compression (see, e.g., Tan 2015)?

Once underway, is cluster formation a fast or slow process relative to the local freefall time (tff)? Tan et al. (2006) and Nakamura & Li (2007) proposed that formation times are relatively long, i.e., ∼10tff, especially for those clusters with high (≳30%) overall star formation efficiency, since simulations of self-gravitating, turbulent, magnetized gas show a low formation efficiency of just ∼2% per freefall time (Krumholz & McKee 2005; Padoan & Nordlund 2011). Alternatively, Elmegreen (2007), Hartmann & Burkert (2007), and Hartmann et al. (2012) have argued for cluster formation in just one or a few freefall times. Another question is: what sets the overall star formation efficiency during cluster formation? The formation timescale and overall efficiency are likely to affect the ability of a cluster to remain gravitationally bound, which on large scales influences global interstellar medium feedback—e.g., concentrated feedback from clusters can create superbubbles (e.g., Krause et al. 2013)—and on small scales controls the feedback environments and tidal perturbations of protoplanetary disks (e.g., Adams 2010).

Star cluster formation is likely to be the result of a complex interaction of numerous physical processes, including turbulence, magnetic fields, and feedback. From the observational side, measuring the structure and kinematic properties of the dense gas component is needed to provide constraints for different theretical models. Previously, Walsh et al. (2004) found small velocity differences between dense cores and surrounding envelopes for a sample of low-mass cores. Kirk et al. (2007, 2010) surveyed the kinematics of over 150 candidate dense cores in the Perseus molecular cloud with pointed ${{\rm{N}}}_{2}{{\rm{H}}}^{+}$ and C18O observations and found subvirial core-to-core velocity dispersions in each region. A similar small core velocity dispersion was also found in the Ophiuchus cloud (André et al. 2007). Qian et al. (2012) searched for 13CO cores in the Taurus molecular cloud and found that the core velocity dispersion exhibits a power-law behavior as a function of the apparent separation, similar to Larson's law for the velocity dispersion of the gas, which suggests that the formation of these cores has been influenced by large-scale turbulence.

These observations have generally focused on nearby low-mass star-forming regions. With the unprecedented sensitivity and spatial resolution of the Atacama Large Millimeter/submillimeter Array (ALMA), more light has been shed on massive star-forming regions from the "clump" scale (of about a few parsecs) to the "core" scale (∼0.01–0.1 pc; e.g., Beuther et al. 2017; Fontani et al. 2018; Lu et al. 2018). Multiple coherent velocity components from filamentary structures have been reported in some massive infrared dark clouds (IRDCs; Henshaw et al. 2013, 2014; Sokolov et al. 2018), similar to the structures seen in the nearby Taurus region by Hacar et al. (2013). "Hub–filament" systems have also been reported in some massive star-forming regions across a variety of evolutionary stages, perhaps indicating the presence of converging flows that channel gas to the junctions where star formation is most active (e.g., Hennemann et al. 2012; Peretto et al. 2014; Lu et al. 2018; Yuan et al. 2018).

However, complete surveys for the dense gas component of massive protoclusters down to the individual core scale are still rare (e.g., Ohashi et al. 2016; Ginsburg et al. 2017), and a large spatial dynamic range is required to perform a multiscale kinematics analysis.

Until recently, only very few nearby regions were known that were candidates for very young and still-forming massive star clusters. One particular promising star-forming clump is G286.21+0.17 (hereafter G286). It is a massive protocluster associated with the η Car GMC at a distance of 2.5 ± 0.3 kpc in the Carina spiral arm (e.g., Barnes et al. 2010). We performed a core mass function study toward this region based on ALMA Cycle 3 observations in Cheng et al. (2018).

Here we present a follow-up study of multiple spectral lines to investigate the gas kinematics and dynamics of G286 from clump to core scales. The paper is organized as follows. In Section 2 we describe the observational setup and analysis methods; the results are presented in Section 3. We discuss the kinematics and dynamics for parsec-scale filaments and dense cores separately in Sections 4 and 5 and then summarize our findings in Section 6.

2. Observational Data

2.1. ALMA Observations

The observations were conducted with ALMA in Cycle 3 (Project ID 2015.1.00357.S; PI: J. C. Tan) during the period from 2015 December to 2016 September. More details of the observations can be found in Cheng et al. (2018). In summary, we divided the region into five strips, denoted as G286_1, G286_2, G286_3, G286_4, and G286_5, each about 1' wide and 5farcm3 long and containing 147 pointings of the 12 m array (see Figure 1). The position of field center is R.A. = 10:38:33, decl. = −58:19:22. We employed the compact configuration C36-1 to recover scales between 1farcs5 and 11farcs0. This is complemented by observations with the ACA, which probes scales up to 18farcs6. Total power (TP) observations were also carried out to recover the total flux (of line emission), which gives a resolution of about 30''.

Figure 1.

Figure 1. Three-color image of G286 constructed by combining Spitzer IRAC 3.6 (blue) and 8.0 (green) μm and Herschel PACS 70 μm (red). Black contours show the 1.3 mm continuum image combining ALMA 12 and 7 m array data (with a resolution of 1farcs62 × 1farcs41). The contour levels are 1σ × (4, 10, 20, 50, 100) with σ = 0.45 mJy beam−1. Gray contours show the 1.3 mm continuum image with only 7 m array data (with a resolution of 7farcs32 × 4farcs42, shown in the lower left corner). The contour levels are 1σ × (4, 10, 20, 50, 100) with σ = 1.7 mJy beam−1. The positions of three filamentary structures are marked in blue text. The G286 field is divided into five strips, as shown by the green rectangles. Each strip is covered with 147 pointings of the 12 m array. The large white ellipse denotes the boundary defined by Mopra HCO+(1–0) emission (Barnes et al. 2011), with the major and minor axes equal to twice the FWHM lengths of the 2D Gaussian fits to its emission. Two smaller ellipses indicate the approximate boundaries of two subregions with distinct radial velocities revealed by multiple gas tracers.

Standard image High-resolution image

During the observations, we set the central frequency of the correlator sidebands to be the rest frequency of the ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$(3–2) line at 231.32 GHz for SPW0 and the ${{\rm{C}}}^{18}{\rm{O}}$(2–1) line at 219.56 GHz for SPW2, with a velocity resolution of 0.046 and 0.048 km s−1, respectively. The second baseband SPW1 was set to 231.00 GHz, i.e., 1.30 mm, to observe the continuum with a total bandwidth of 2.0 GHz, which also covers CO(2–1) with a velocity resolution of 0.64 km s−1. The frequency coverage for SPW3 ranges from 215.85 to 217.54 GHz to observe DCN(3–2), DCO+(3–2), SiO(v = 0)(5–4), and ${\mathrm{CH}}_{3}\mathrm{OH}({5}_{\mathrm{1,4}}-{4}_{\mathrm{2,2}})$. This paper will focus mostly on dense gas tracers C18O, ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$, ${\mathrm{DCO}}^{+}$, and $\mathrm{DCN}$.

The raw data were calibrated with the data reduction pipeline using CASA 4.7.0. The continuum visibility data were constructed with all line-free channels. We performed imaging with the tclean task in CASA, and during cleaning, we combined data for all five strips to generate a final mosaic map. Two sets of images were produced for different aspects of the analysis, one including the TP and 7 m array data and one combining the TP and 7 and 12 m data. The 7 m array data were imaged using a Briggs weighting scheme with a robust parameter of 0.5, which yields a resolution of 7farcs32 × 4farcs42. For the combined data, we used the same Briggs parameter. In addition, since we have extra uv coverage for part of the data, we also apply a 0farcs6 uvtaper to suppress longer baselines, which results in 1farcs62 × 1farcs41 resolution. Both image sets are then feathered with the TP image to correct for the missing large-scale structures. Our sensitivity level is about 30 mJy beam–1 per 0.1 km s−1 for ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$ and ${{\rm{C}}}^{18}{\rm{O}}$. A sensitivity of 45 mJy beam–1 per 0.1 km s−1 is achieved for ${\mathrm{DCO}}^{+}$(3–2), $\mathrm{DCN}$(3–2), SiO(5–4), and ${\mathrm{CH}}_{3}\mathrm{OH}$(51,4 − 42,2).

2.2. Herschel Observations

The far-IR dust continuum images of G286 were taken from the Herschel Infrared GALactic plane survey (Molinari et al. 2010, 2016). The data include Photodetector Array Camera and Spectrometer (PACS; 70 and 160 μm) and Spectral and Photometric Imaging REceiver (250, 350, and 500 μm) images. We performed pixel-by-pixel graybody fits to derive the mass surface density (Σ) of the G286 region, following the procedures in Lim et al. (2016). The background was estimated as the median intensity value between two and four times the ellipse aperture shown in Figure 1. To better probe the smaller, higher Σ structures, we generated a higher-resolution Σ map by regridding the λ ∼ 160–500 μm images to match the 250 μm data (see Lim et al. 2016, for details).

3. General Results

An overview of the observed region and the layout of the ALMA observations are shown in Figure 1. With the large spatial dynamic range of the ALMA data set, we will present the large-scale structures traced with single-dish TP observations first, followed by higher-resolution 7 and 12 m array observations.

3.1. Observations with the TP Array

Figure 2(a) shows the spectra of CO(2–1) and C18O(2–1) averaged inside a 2farcm5 radius aperture centered on the phase center. The CO(2–1) line has a maximum around the known systemic velocity of about −20 km s−1. A secondary, much weaker peak is seen around −9 km s−1. This component is also seen in the Mopra CO(2–1) map, which appears to be a diffuse structure larger than our field of view. We expect that this feature is probably contributed by a foreground or background cloud along the line of sight, and there is no indication of an interaction between this cloud and G286. Emission from C18O(2–1) is only seen from the main −20 km s−1 component.

Figure 2.

Figure 2. (a) Averaged CO(2–1) and C18O(2–1) TP spectra extracted over a 2farcm5 radius aperture centered on the phase center. Note that the flux scale of CO(2–1) has been reduced by a factor of 10. (b) Same as (a) but for C18O(2–1), ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$(3–2), and ${\mathrm{DCO}}^{+}$(3–2) in a smaller velocity range from −30 to −10 km s−1. The flux scale of C18O(2–1) is reduced by a factor of 20 and that of ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$(3–2) is increased by a factor of 3 for ease of comparison. Note that the ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$(3–2) emission is affected by hyperfine structure, while ${\mathrm{DCO}}^{+}$(3–2) is not.

Standard image High-resolution image

Figure 2(b) shows the spectra of the deuterated dense gas tracer ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$(3–2) and ${\mathrm{DCO}}^{+}$(3–2), averaged over the same region and compared to C18O(2–1), zooming in to the velocity range of the main −20 km s−1 component. Deuterated species, such as ${\mathrm{DCO}}^{+}$ and ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$, are expected to be tracers of cold, dense gas, including material that is contained in prestellar cores (e.g., Crapsi et al. 2005; Bergin & Tafalla 2007; Kong et al. 2015), and typically optically thin even at the core scale, as found in some examples in IRDCs (Tan et al. 2013). Interestingly, the C18O(2–1) line exhibits a main Gaussian-like profile with a slight skewness (or second component) to the redshifted side. The spectrum from ${\mathrm{DCO}}^{+}$(3–2) shows a more pronounced double-peaked profile, with one component at about −20.5 km s−1 and the other at −18.5 km s−1.

The double-peak profile, i.e., with a stronger blue wing, has also been seen in the ${\mathrm{HCO}}^{+}$(1–0) and ${\mathrm{HCO}}^{+}$(4–3) lines in Barnes et al. (2010), with similar central velocities for both peaks. It was interpreted by Barnes et al. (2010) as a canonical inverse P Cygni profile indicating gravitational infall (Zhou et al. 1993). However, in this picture, we would expect a single Gaussian profile for optically thin tracers at the self-absorption velocity, in contrast to our ${\mathrm{DCO}}^{+}$(3–2) spectrum. We will return in Section 5 to the question of whether the claimed inverse P Cygni profile in ${\mathrm{HCO}}^{+}$ is really tracing global clump infall or whether it is arising from distinct spatial and kinematic substructures in the protocluster.

To further explore the kinematic structure of the clump, we present the CO(2–1) channel map from −55.0 to 15.0 km s−1 in Figure 3(a), where we have averaged four velocity channels in each displayed panel. The CO emission is widespread around the systemic velocity (−23 to −17 km s−1). Blueward of the line center, the emission retains extension toward the SE and then, at the highest blueshifted velocities, e.g., v ≲ −45 km s−1, appears more concentrated. The redshifted emission shows more complex structure, including from emission features already mentioned at around v = −9 km s−1, which may be from an unrelated cloud along the line of sight. However, high-velocity (Δv ≳ 25 km s−1) redshifted gas is still seen near the phase center. These high-velocity features, both blue- and redshifted, are likely to be caused by protostellar outflow activity from within the G286 star-forming clump.

Figure 3.

Figure 3. (a) Channel maps of TP CO(2–1) emission integrated over every 2.0 km s−1, as indicated in the upper left corner of each panel (indicating the central velocity of the range), from −55.0 to +15.0 km s−1. The contour levels are 1 Jy beam−1 km s−1 × (1, 5, 10, 20, 40, 80, 200). The red plus sign in each panel marks the phase center of the observation (R.A. = 10:38:33, decl. = −58:19:22). The thick black contour in the lower left panel shows the 4σ level of the 7 m continuum emission. (b) Channel maps of TP C18O(2–1) emission integrated over every 1.0 km s−1, with ranges from −22.5 to −17.5 km s−1. The contour levels are 1 Jy beam−1 km s−1 × (1, 5, 10, 20, 40, 80, 200). The thick black contour in the left panel shows the 4σ level of the 7 m continuum emission.

Standard image High-resolution image

The clump-averaged spectra could be affected by multiple factors, including collapse, rotation, and outflows. To better resolve the kinematics near the systemic velocity, where 12CO(2–1) is expected to be mostly optically thick, in Figure 3(b), we show the C18O(2–1) channel map from −23.0 to −17.0 km s−1. This C18O emission at around −20 km s−1 is moderately elongated in the N–S direction. In the central 2' region, the C18O(2–1) at blueshifted velocities is mostly extended to the SE, while at the corresponding redshifted velocities, there is a more complex, widespread morphology, including some material at NE and SE locations.

3.2. Observations of the 7 and 12 m Arrays

Figure 4 presents summary maps of four spectral lines, C18O(2–1), ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$(3–2), ${\mathrm{DCO}}^{+}$(3–2), and $\mathrm{DCN}$(3–2) (left to right), overlaid on a 1.3 mm continuum image in black contours. The top two rows show the moment 0 and moment 1 maps of the 7 m array images, respectively. As shown in the moment 0 map, C18O traces structures that are more spatially extended than other lines. Here ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$ and ${\mathrm{DCO}}^{+}$ are more closely associated with the dust continuum, but their distributions are slightly different; ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$ is mainly detected toward the NW–SE filament and the southern part of the NE–SW filament. Note also that not all of the regions with strong dust continuum have detections of ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$. In particular, there is a deficiency of ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$ emission toward the central brightest clump. On the other hand, ${\mathrm{DCO}}^{+}$ emission appears slightly more extended. There is also an E–W filamentary feature to the E of the NW–SE filament. This E–W filament is not seen clearly in continuum emission, where we only observe a few cores strung out along the E–W direction, but these do appear to be connected by weak diffuse dust emission seen at a 3σ level. Additionally, ${\mathrm{DCO}}^{+}$ is also detected toward a few positions to the S of the NW–SE filament. The spatial distribution of $\mathrm{DCN}$ emission is dramatically different from that of ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$ and ${\mathrm{DCO}}^{+}$: it is strongly concentrated toward the clump in the center, where no detection or only weak detection is seen for ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$ and ${\mathrm{DCO}}^{+}$. The $\mathrm{DCN}$ emission becomes weaker away from the center.

Figure 4.

Figure 4. Summary figure for the 7 and 12 m line observations. Columns from left to right show the results of C18O(2–1), ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$(3–2), ${\mathrm{DCO}}^{+}$(3–2), and $\mathrm{DCN}$(3–2). From top to bottom, the color scales show the maps of 7 m moment 0, 7 m moment 1, 7 m+12 m moment 0, and 7 m+12 m moment 1. The color bar on the right indicates the flux scale in Jy beam−1 for moment 0 maps and velocity in km s−1 for moment 1 maps. The black contours illustrate the 1.3 mm continuum emission for comparision, with the first two rows showing 7 m continuum images and the last two rows 7 m+12 m images.

Standard image High-resolution image

The different morphological distributions of the deuterated species may be due to chemical differentiation. In general, ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$ is known as a good tracer of cold (T ≲ 20 K), dense gas, where ${{\rm{H}}}_{2}{{\rm{D}}}^{+}$ builds up in abundance but CO is mostly frozen out onto dust grains (e.g., Fontani et al. 2015). Formation of ${\mathrm{DCO}}^{+}$ requires both ${{\rm{H}}}_{2}{{\rm{D}}}^{+}$ and gas phase CO (e.g., Millar et al. 1989), which requires a temperature ≲30 K but not too cold to cause significant CO freeze-out. On the other hand, the primary DCN formation mechanisms are thought to require ${\mathrm{CH}}_{2}{{\rm{D}}}^{+}$ instead of ${{\rm{H}}}_{2}{{\rm{D}}}^{+}$, which is energetically favorable up to ∼80 K (e.g., Millar et al. 1989; Turner 2001). Additionally, sputtering from grain mantles can also lead to enhancement of the DCN abundance in shocked regions (e.g., Busquet et al. 2017). Hence, we would generally expect more DCN emission in relatively later evolutionary stages. The concentrated distribution of DCN, combined with more widespread ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$ and ${\mathrm{DCO}}^{+}$ emission, indicates a scenario that star formation, especially more massive, luminous star formation, has taken place first in the central regions of G286 compared to in the more extended filaments.

The second row of Figure 4 shows the moment 1 map of the 7 m array images. The C18O moment 1 map reveals redshifted emission associated with the NE–SW filament and then continuing to the S of the NW–SE filament, while the NW–SE and E–W filaments are mainly associated with blueshifted gas. Other dense gas tracers show similar velocity patterns as C18O but with the emission mainly detected toward dense continuum clumps. In particular, DCN illustrates the blue–red velocity transition across the central clump in the NW–SE direction.

A zoom-in view of G286 is presented in the third and fourth rows of Figure 4, illustrating the moment 0 and moment 1 maps of the combined 12 m+7 m array image with a resolution of ∼1farcs5. The continuum image reveals a higher level of fragmentation and many well-defined dense cores, with a typical size of a few thousand au. The E–W filament and part of the NE–SW filament are resolved out in this continuum image. The intensity and velocity distribution of C18O appears more complicated seen in high resolution. Other dense gas tracers, like ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$, still have good association with the continuum at the core scale, and the velocity pattern is also consistent with that seen in the 7 m image.

In Figure 5, we present the 12 m+7 m C18O image with integrated emission in different velocity intervals shown in different colors, i.e., −23.0 to −20.5 km s−1 in blue, −20.5 to −19.5 km s−1 in green, and −19.5 to −17.0 km s−1 in red. Besides the velocity structures seen in Figure 4, this plot also reveals highly filamentary C18O features around the systemic velocity. These filaments are more spatially extended than the continuum. While some of this morphology may be affected by artificial side lobes from imperfect cleaning of the interferometric data, at least some of the C18O(2–1) filaments have corresponding detections in the continuum and hence are most likely to be real features.

Figure 5.

Figure 5. Three-color image constructed with integrated 12 m+7 m C18O(2–1) emission (−23.0 to 20.5 km s−1 in blue, −20.5 to −19.5 km s−1 in green, and −19.5 to 17.0 km s−1 in red). The synthesized beam (1farcs56 × 1farcs40) is shown in the lower left corner. The 7 m continuum image is shown in white contours for comparison. The contour levels are 1.7 mJy beam−1 × (3, 6, 10, 20, 50, 100).

Standard image High-resolution image

4. Filamentary Virial Analysis

As shown in Figure 1, the millimeter continuum emission reveals two main filaments: a northern one with an NE–SW orientation and a southern one with an NW–SE orientation. Here we perform a filamentary virial analysis following Fiege & Pudritz (2000). Since the NE–SW filament is mostly filtered out at higher resolution, we utilize the 7 m array data (continuum and C18O) for this section.

Figure 6 illustrates the C18O(2–1) emission integrated over every 0.5 km s−1 from −22.0 to −17.0 km s−1. As in Figure 5, filamentary structures are seen near the systemic velocity of −20 km s−1. At least three of the C18O(2–1) filaments have corresponding detections in the continuum at a 2σ level and hence are most likely real features, rather than side-lobe artifacts. The most prominent filament is associated with the NE–SW continuum filament and is clearly seen from −20.0 to −17.0 km s−1. The NW–SE filament appears more complicated in C18O(2–1) and is not well described as being a coherent C18O filamentary structure. Therefore, we carry out a virial analysis only for the NE–SW filament.

Figure 6.

Figure 6. The 7 m C18O(2–1) emission integrated over 0.5 km s−1 intervals, as indicated in the upper left corner of each panel, from −22.0 to −17.0 km s−1. The black contours show the 7 m array 1.3 mm continuum emission. The contour levels are 1.7 mJy beam−1 × (4, 10, 20, 50, 100).

Standard image High-resolution image

As shown by Fiege & Pudritz (2000), a pressure-confined, nonrotating, self-gravitating, filamentary (i.e., length ≫ width) magnetized cloud that is in virial equilibrium satisfies

Equation (1)

where Pf is the mean total pressure in the filament, Pe is the external pressure at its surface, mf is its mass per unit length, ${m}_{\mathrm{vir},{\rm{f}}}=2{\sigma }_{f}^{2}/G$ is its virial mass per unit length, and Mf and Wf are the gravitational energy and magnetic energy per unit length, respectively. Here, because of the observational difficulties of measuring the surface pressure and magnetic fields, we ignore the surface and magnetic energy terms, i.e., only considering the balance between gravity and internal pressure support.

To measure the properties of the filament, we show in Figure 7(a) a 60'' × 20'' rectangle that closely encompasses the NE–SW filament, which we use to define the filament boundary. From the Herschel spectral energy distribution (SED)–derived mass surface density map, we find average values of Σsed in the strips ranging from 0.25 (in strip 1, which is closest to the center of G286) to 0.12 g cm−2 (in strip 4; see Table 1). The mass in each region is then estimated, with values between MSED = 26 and 53 M. For comparison, we also calculate masses from the 1.3 mm continuum flux, assuming a temperature of 20 K and other dust properties following Cheng et al. (2018). We find that the 1.3 mm–derived mass estimates are about a factor of 2 smaller than that measured from the Herschel SED fitting method. Since the ALMA 7 m array observations only probe scales up to ∼19'', they are likely to be missing some flux from the filament leading to an underestimation of the masses, and so here we adopt the Herschel SED-derived mass estimates for the virial analysis.

Figure 7.

Figure 7. (a) Column density map made with Herchel submillimeter continuum data, overlaid on the 7 m array continuum emission in contours. The contour levels are 1.7 mJy beam−1 × (4, 10, 20, 50, 100). The ALMA synthesized beam is shown in the lower left corner, while the resolution of the Herschel-derived mass surface density map is shown in the lower right corner. The red rectangles dilineate the position of the NE–SW filament and its division into four strips, numbered 1–4 from S to N. (b) C18O(2–1) spectra of the four strips of the NE–SW filament and the total (see legend). The green lines show primary Gaussian component fits to these spectra.

Standard image High-resolution image

Table 1.  Properties of the NE–SW Filament

Properties Strip 1 Strip 2 Strip 3 Strip 4 Total
${\overline{{\rm{\Sigma }}}}_{\mathrm{sed}}$ (g cm−2) 0.25 0.17 0.13 0.12 0.17
MSED (M) 53 36 27 26 142
M1.3mm (M) 25 17 21 11 74
mSED,f (M pc−1) 254 170 130 123 170
${\overline{v}}_{f}$ (km s−1) −17.85 −18.59 −19.01 −19.40 −18.73
${\sigma }_{{{\rm{C}}}^{18}{\rm{O}}}$ (km s−1) 0.40 0.52 0.53 0.61 0.52a
σf (km s−1) 0.48 0.58 0.59 0.66 0.58
mvir,f (M pc−1) 106 158 160 204 158
mf/mvir,f 2.39 1.08 0.81 0.60 1.08

Note.

aFor velocity dispersion, we take the linear average of four strips.

Download table as:  ASCIITypeset image

The 60'' length of the filament corresponds to 0.73 pc at an assumed distance of 2.5 kpc. We assume a 10% uncertainty in the distance (e.g., Barnes et al. 2010). Without direct observational constraints, we further assume that the filament axis is inclined by an angle i = 60° to the line of sight (90° would be in the plane of the sky). If an inclination angle of 90° or 30° were to be adopted, then the length estimates would differ by factors of 1.15 and 0.577, respectively. Thus, the actual length of the filament is assumed to be 0.84 pc (or 3/4 of this from the centers of strip 1 to strip 4). Thus, the overall mass per unit length of the filament is mSED,f ∼ 170 M pc−1, with strip 1 having a higher value of ∼250 M pc−1.

The mean line-of-sight velocity and velocity dispersion of the filament are measured from the average C18O spectra inside the rectangular regions. To reduce contamination from surrounding ambient gas at the systemic velocity, we utilize the image cube made with only the 7 m array data (i.e., without feathering with the TP data), as illustrated in Figure 7(b). We perform Gaussian fitting to measure the average centroid velocity ${\overline{v}}_{f}$ and velocity dispersion ${\sigma }_{{{\rm{C}}}^{18}{\rm{O}}}$. For strips 1 and 3, two Gaussian components are used for a better fitting. We then compare the C18O spectra with the spectra of denser gas tracers like ${\mathrm{DCO}}^{+}$ and find that only one component is associated with these tracers. This primary component is shown by green lines in Figure 7(b) and is used for further analysis.

The values of ${\overline{v}}_{f}$ show a steady progression from −17.85 km s−1 in strip 1 to −19.40 km s−1 in strip 4, which corresponds to an overall velocity gradient of 2.84 km s−1 pc−1 using the plane-of-sky projected distance or 2.46 km s−1 pc−1 for the assumed 60° inclination. We can compare these kinematics to the IRDC filament studied by Hernandez et al. (2011, 2012), which has a length of 3.77 pc on the sky (4.35 pc for the assumed 60° inclination) and also had its C18O(2–1) emission analyzed in four strips. Here the velocities did not show a steady progression but showed differences of about 0.5 km s−1 from strip to strip, i.e., corresponding to velocity gradients of about 0.53 km s−1 pc−1 in the plane of the sky. The larger and more systematic velocity gradient shown in the NE–SW filament in G286 may be the result of acceleration due to infall into the protocluster potential. Strip 4 has a mean velocity similar to that of the ambient, larger-scale gas in the region, while strip 1, closer in projection to the protocluster center, is redshifted with respect to this velocity. Thus, in this scenario, the strip 4 end of the filament is closer to us than the protocluster center.

If the velocity change from strip 4 to strip 1, i.e., +1.55 km s−1, is due to infall in the protocluster potential, then we can use this information to constrain the mass of the protocluster. Assuming a uniform distribution of matter in a spherical protocluster clump of radius L, the change in potential from the edge to the center is GM/(2L). If material starts at rest at a radius L, i.e., the strip 4 position, and then accelerates to a velocity v1, of which we observe v1 cos i, then the mass inside the radius L is

Equation (2)

For an observed length Lobs from the center of strip 4 to the center of strip 1 of 0.55 pc (i.e., 3/4 of 0.73 pc) and a line-of-sight velocity difference of 1.55 km s−1, we thus estimate the dynamical mass to be 1410 M, assuming i = 60°. If an inclination angle of 30° or 70° is adopted, the mass would be 814 or 2780 M, respectively. This estimation based on filament infall kinematics is broadly consistent with that derived from Herschel SED fitting, i.e., ∼2900 M for the region defined by the larger ellipse aperture in Figure 1. Note that for this SED-based method, we expect ∼50% uncertainty in the mass estimation due to dust opacity and distance uncertainties.

Considering the internal dynamics of the filament, in order to account for support against gravity from both thermal and nonthermal motions of the gas, we subtract the thermal component of broadening of the C18O(2–1) line from the measured velocity dispersion (in quadrature, assuming a temperature of 20 K) and add back the sound speed to obtain the total 1D velocity dispersion, σf, i.e.,

Equation (3)

where μp = 2.33 is the mean molecular weight assuming nHe = 0.1nH and ${\mu }_{{{\rm{C}}}^{18}{\rm{O}}}$ is the molecular weight of C18O. We have then carried out a virial analysis for each of the four strips (see Table 1). Note that for strips 1 and 3, we fit the spectra with two Gaussian components and utilize the component that is more clearly associated with the filament. For example, in strip 3, the velocity component near −20.5 km s−1 is contributed by another gas clump to the NW of the filament.

The values of mf/mvir,f of the four strips range from 0.60 to 2.39. Given the systematic uncertainties in measuring the masses and lengths of the structures that combine to be at least ∼50%, these values are consistent with the filament being in approximate virial equilibrium, even without accounting for surface pressure and magnetic support terms. We also note that the values of mf/mvir,f grow, i.e., become less gravitationally bound, as one progresses from strip 4 to strip 1. This may indicate that infall motions and/or tidal forces toward the center of the protocluster act to stabilize the filament.

5. Kinematic Properties of the Dense Core Sample

Cheng et al. (2018) analyzed the mass distribution of dense cores toward the central region of G286 (about 2farcm× 1farcm5), where the uv coverage of the observation allows imaging with ∼1'' resolution. Here we carry out a kinematic follow-up study on the dense core sample in this region.

Figure 8 shows the integrated intensity map of C18O(2–1), ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$(3–2), ${\mathrm{DCO}}^{+}$(3–2), and $\mathrm{DCN}$(3–2) in the central region, with three velocity ranges shown in different colors. This map is similar to the 12 m+7 m moment maps in Figure 4 but emphasizes relatively weaker features that might be missing in Figure 4 due to higher noise resulting from its wider velocity range. Most cores in this region have significant detection from at least one of the three dense gas tracers, ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$(3–2), ${\mathrm{DCO}}^{+}$(3–2), and $\mathrm{DCN}$(3–2), and this allows us to measure the centroid velocity and velocity dispersion for each dense core.

Figure 8.

Figure 8. (a) C18O integrated intensity map using combined 7 m+12 m array data. Red, green, and blue contours show emission integrated from −23 to −21, −21 to 19, and −19 to −17 km s−1, respectively. The contours start from 4σ in steps of 2σ, with σ = 0.1 Jy beam−1 km s−1. The gray-scale image is the 1farcs0 resolution 7 m+12 m array combined 1.3 mm continuum image. (b) Same as panel (a) but for ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$(3–2). The contours start from 4σ in steps of 2σ, with σ = 0.025 Jy beam−1 km s−1. (c) Same as panel (a) but for ${\mathrm{DCO}}^{+}$(3–2). The contours start from 4σ in steps of 2σ, with σ = 0.03 Jy beam−1 km s−1. (d) Same as panel (a) but for $\mathrm{DCN}$(3–2). The contours start from 4σ in steps of 2σ, with σ = 0.03 Jy beam−1 km s−1.

Standard image High-resolution image

5.1. Review of the Core Sample Based on Dust Continuum Emission

Cheng et al. (2018) reported different numbers of identified cores, ranging from 60 to 125, depending on the detection algorithm and parameter choices of these algorithms. Here we adopt the fiducial dendrogram-identified core sample with a base threshold of 4σ and a delta threshold of 1σ, along with a minimum area of half a synthesized beam size. This parameter combination yields 76 cores.

In Appendix A, we list the properties of the dense core sample. The cores are here named as G286c1, G286c2, etc., with the numbering order from highest to lowest core mass. The masses are estimated to range from 0.19 to 80 M, assuming a constant temperature of 20 K for each core (see Cheng et al. 2018 for more details). The radius is evaluated as Rc = $\sqrt{A/\pi }$, where A is the projected area of the core. The median radius is 0.011 pc, similar to the spatial resolution (∼1'', 2500 au), indicating that many cores are not well resolved. Note that we adopt the core area returned by dendrogram, which is defined with an isophotal boundary at a certain flux level, i.e., the level where two cores merge together or the 4σ flux threshold for isolated cores. So the core area or radius could be underestimated in a crowded field.

We then evaluate the mean mass surface density of the cores as Σc ≡ M/A. The median mass surface density of our sample is ∼0.65 g cm−2, and all cores have values ≳0.4 g cm−2. We also evaluate the mean H nuclei number density in the cores, nH,c ≡ Mc/(μHV), where μH = 1.4mH is the mean mass per H assuming nHe = 0.1nH and V = $4\pi {R}_{c}^{3}/3$. The mean value of log10(nH,c/cm−3) is 6.88, with a standard deviation of 0.24.

5.2. Spectral Fitting

We extract the average C18O(2–1), ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$(3–2), ${\mathrm{DCO}}^{+}$(3–2), and $\mathrm{DCN}$(3–2) spectra of each core, which are shown in Appendix B. Among the four tracers, C18O is the strongest for almost all cores, and sometimes the C18O profiles can be complex. Other lines are relatively weak and only detected for part of the core sample.

To measure the centroid velocity and velocity dispersion of each core, we only fit spectra with well-defined profiles, i.e., those with a peak greater than a certain threshold value. Here we adopt a 4σ criterion for this threshold value. Since the noise levels of the average spectra vary for different cores (depending on the pixel numbers in the core, etc.), we estimate the rms noise separately for each core and line using the signal-free channels. This signal-to-noise criterion gives 74 cores detected in C18O(2–1) (97%), 27 in ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$(3–2) (36%), 45 in ${\mathrm{DCO}}^{+}$(3–2) (59%), and 29 in $\mathrm{DCN}$(3–2) (38%). We also checked the single pixel spectra at the continuum peak of each core and found that the vast majority have similar line profiles as the averaged spectra, but the signal-to-noise ratios are usually lower, so we proceed with our analysis using the core-averaged spectra.

We characterize the C18O(2–1) spectra with 1D Gaussian fitting using the curve_fit function in the Scipy.optimize Python module. Most cores can be well described with a single Gaussian component. In general, we expect that C18O(2–1) traces lower-density envelope gas surrounding the dense core and thus could be more affected by multiple components along the line of sight. In 31 cores where a spectrum has more complex profiles and hence cannot be well approximated by a single Gaussian, we allow for a second Gaussian component.

For the ${\mathrm{DCO}}^{+}$(3–2) and $\mathrm{DCN}$(3–2) lines, we also perform the Gaussian fitting with the curve_fit function. For the ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$(3–2) line, to account for the full blended hyperfine components, we use the hyperfine line-fitting routine in pyspeckit (Ginsburg & Mirocha 2011), with the relative frequencies and optical depths for ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$ taken from Dore et al. (2004) and Pagani et al. (2009). These dense gas tracers are usually well described with one Gaussian component. Figure 9 shows a example of the line fitting. In particular, in one case (i.e., G286c3), two separate components were clearly required for a good fit for ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$ and ${\mathrm{DCO}}^{+}$. These two components are mostly likely to belong to two separate entities that are not resolved in their continuum emission.

Figure 9.

Figure 9. Example of the spectral line fitting for G286c5. Here we use one Gaussian component to fit the spectra of ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$(3–2), ${\mathrm{DCO}}^{+}$(3–2), and $\mathrm{DCN}$(3–2) and two components for C18O(2–1).

Standard image High-resolution image

5.3. Comments on Individual Cores

G286c1. This is the most massive core in G286, with a mass of 80 M and an equivalent radius of 0.036 pc. It is associated with strong infrared emission and a wide angle bipolar CO outflow (Y. Cheng et al. 2020, in preparation), and hence it is already in a relatively evolved protostellar stage. If we adopt a higher temperature, such as 70 K, typical of massive protostellar sources (e.g., Zhang & Tan 2018), then its mass would be ∼20 M. It is not detected in ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$(3–2), but we see broad line profiles from ${\mathrm{DCO}}^{+}$(3–2), $\mathrm{DCN}$(3–2), and C18O(2–1). In particular, there is very strong $\mathrm{DCN}$(3–2) emission from −22 to −15 km s−1, which is even broader than C18O. Our high-resolution ALMA observation in Cycle 5 has revealed further fragmentation and substructures in G286c1 (Y. Cheng et al. 2020, in preparation). Here we still use a one Gaussian component to model the spectral lines of G286c1, and the resulting fitting parameters should be treated more cautiously as reflecting the average properties of the core.

G286c3. This is also a massive core, with ∼12 M. We have detected C18O(2–1), ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$(3–2), ${\mathrm{DCO}}^{+}$(3–2), and $\mathrm{DCN}$(3–2) toward G286c3. Interestingly, these spectra of C18O, ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$, and ${\mathrm{DCO}}^{+}$ all exhibit a double-peak profile, with one peak centered around −19.5 km s−1 and another at ∼−18 km s−1, though $\mathrm{DCN}$ is only detected in one velocity component. Since we expect the deuteratated species, such as ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$ and ${\mathrm{DCO}}^{+}$, to be optically thin, these line profiles are more likely to be contributed by two separate entities inside G286c3 instead of a central dip caused by self-absorption. A detailed inspection from the continuum also reveals that G286c3 is very elongated in the NE–SW direction. Thus, it is possible that there are further subfragmentations in G286c3 that are not identified by our fiducial dendrogram algorithm; e.g., there could be two cores overlapping along the line of sight. Here we use two-component Gaussian fitting to model the spectrum of C18O, ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$, and ${\mathrm{DCO}}^{+}$ and treat them as two individual cores (i.e., two data points per line in Figure 10). We split the mass of G286c3 assuming that the mass of each component is proportional to the C18O flux for relevant analysis.

Figure 10.

Figure 10. Centroid core velocity and velocity dispersion of each core measured with C18O(2–1), ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$(3–2), ${\mathrm{DCO}}^{+}$(3–2), and $\mathrm{DCN}$(3–2).

Standard image High-resolution image

G286c4. This core has an estimated mass of ∼9 M. There is no ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$ detection, but we see very strong C18O, ${\mathrm{DCO}}^{+}$, and $\mathrm{DCN}$ emission. Here $\mathrm{DCN}$(3–2) has a very strong peak centered at −19.5 km s−1, similar to C18O and ${\mathrm{DCO}}^{+}$. Additionally, there are two secondary peaks at around velocities of −16 and −23 km s−1. These may be real features resulting from unresolved condensations or more dynamical activities like outflows, but we are unsure about their origin with the current information. Here for $\mathrm{DCN}$, we only fit the central major velocity component that is consistent with other tracers.

G286c8, G286c20, and G286c41. These are special in terms of their ${\mathrm{DCO}}^{+}$(3–2) spectra. All three cores have a ${\mathrm{DCO}}^{+}$ peak around −18 km s−1. For G286c20 and G286c41, ${\mathrm{DCO}}^{+}$ has a large velocity offset (∼1 km s−1) compared with other tracers, like C18O. For G286c8, this offset is even larger (∼3 km s−1), and there is another obvious ${\mathrm{DCO}}^{+}$ peak around −21 km s−1, similar to the peaks of the C18O and $\mathrm{DCN}$ lines. A possible explanation is that G286c8 has a core velocity around −21 km s−1, as traced by multiple tracers, while the ${\mathrm{DCO}}^{+}$ feature around −18 km s−1 is not associated with the dense core. From the continuum map, we find that all three cores are close together and lie on a filamentary feature that is only seen in ${\mathrm{DCO}}^{+}$. This filamentary feature is clear in the ${\mathrm{DCO}}^{+}$ channel map and does not appear to be associated with a dense dust continuum. Hence, we exclude this ${\mathrm{DCO}}^{+}$ velocity component near −18 km s−1 for G286c8, G286c20, and G286c41 in our analysis.

5.4. Line Parameters of Different Tracers

The best-fit parameters of centroid velocity and velocity dispersion are displayed along with the spectral lines in Appendix B. Figure 10 illustrates the distribution of these parameters, along with their individual uncertainties. As can be seen, the centroid velocities range from −22.5 to −17.0 km s−1, and there is a modest clustering near −21.5 km s−1. The velocity dispersions range from 0.1 to 1.0 km s−1 for deuterated species, while those of C18O are systematically larger. Here ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$ and ${\mathrm{DCO}}^{+}$ usually give smaller velocity dispersions, with median values of 0.35 and 0.36 km s−1, respectively. The $\mathrm{DCN}$-measured dispersions are larger, with a median value of 0.43 km s−1. Centroid velocity uncertainties range from 0.01 to 0.08 km s−1, while velocity dispersion uncertainties vary from 1% to 20%, with a few cases ≳30%, depending on the signal-to-noise ratio and shape of the line profiles.

Figure 11 illustrates the line detection situation of the core sample, with different colored circles denoting detections in ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$(3–2), ${\mathrm{DCO}}^{+}$(3–2), and $\mathrm{DCN}$(3–2). As C18O(2–1) is detected for almost all of the cores (except c68 and c75), it is not shown here. As already apparent in Figure 8, $\mathrm{DCN}$(3–2) is mostly detected in the central region, while ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$(3–2)– and ${\mathrm{DCO}}^{+}$(3–2)–detected cores are more widespread. Overall, we have 54 cores that are detected in at least one of these three dense gas tracers. In particular, all of the cores with ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$ detection also have strong ${\mathrm{DCO}}^{+}$ emission.

Figure 11.

Figure 11. Left: line detection status for each core overlaid on the 1farcs0 resolution 1.3 mm continuum image in contours. The black plus signs denote the positions of cores identified via 1.3 mm continuum by Cheng et al. (2018). A red circle indicates a detection of ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$(3–2), a blue circle ${\mathrm{DCO}}^{+}$(3–2), and a green circle $\mathrm{DCN}$(3–2). Right: core velocity map overlaid on the 1farcs0 resolution 1.3 mm continuum image shown in contours and gray scale. The core velocity is determined by averaging the results from ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$, ${\mathrm{DCO}}^{+}$, and $\mathrm{DCN}$ (see text). The colored circles indicate the velocities of the 54 dense cores that are detected in at least one deuterated tracer. Overall, the velocity distribution can be described as being composed of two velocity groups that are spatially distinct, with an approximate boundary shown by the dashed green line.

Standard image High-resolution image

The cores that are detected in more than one line are of particular interest, since differences in fitted parameters could be a reflection of chemical differentiation. There are 14 cores that are detected in all three lines, 26 in both ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$ and ${\mathrm{DCO}}^{+}$, 14 in both ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$ and $\mathrm{DCN}$, and 21 in both ${\mathrm{DCO}}^{+}$ and $\mathrm{DCN}$. Figure 12 illustrates the differences in the fitted parameters of these species when commonly detected. From this figure we see that there is no significant offset in centroid velocity or velocity dispersion as derived from the different species. This similarity in velocity distributions is expected if these species are tracing the same molecular gas material.

Figure 12.

Figure 12. Core centroid velocity differences and relative velocity dispersions as measured from ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$(3–2), ${\mathrm{DCO}}^{+}$(3–2), and DCN(3–2).

Standard image High-resolution image

For the centroid velocities, the median offsets between ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$ and ${\mathrm{DCO}}^{+}$, ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$ and $\mathrm{DCN}$, and ${\mathrm{DCO}}^{+}$ and $\mathrm{DCN}$ are 0.07, 0.09, and 0.03 km s−1, respectively. The sampling error of the velocity offset distribution due to the finite number of cores is estimated to be about 0.04 km s−1, so these offsets are not very significant.

The 1D velocity dispersion σ are generally consistent among different tracers within a factor of 2. The median values of ${\sigma }_{\mathrm{DCN}}/{\sigma }_{{\mathrm{DCO}}^{+}}$, ${\sigma }_{{{\rm{N}}}_{2}{{\rm{D}}}^{+}}/{\sigma }_{\mathrm{DCN}}$, and ${\sigma }_{{{\rm{N}}}_{2}{{\rm{D}}}^{+}}/{\sigma }_{{\mathrm{DCO}}^{+}}$ are 1.16, 0.99, and 0.95, respectively. The observed scatter is consistent with the fitting uncertainties.

We next compare dense core centroid velocities with the larger-scale gas reservoir (or envelope) traced by C18O. Previous studies in relatively low-mass environments have shown that cores mostly have subsonic core-to-envelope motions (e.g., Walsh et al. 2004, 2007; Kirk et al. 2007; Walker-Smith et al. 2013). Our work here provides a measure of core-to-envelope motions within a more massive protocluster. Additionally, most previous works measured the centroid velocity offset between C18O and ${{\rm{N}}}_{2}{{\rm{H}}}^{+}$. Here we have observations of lines from deuterated species like ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$, ${\mathrm{DCO}}^{+}$, and $\mathrm{DCN}$, which should be better tracers of localized dense cores rather than more extended filaments and usually not affected by multiple velocity components that may complicate the interpretation (e.g., Henshaw et al. 2014; Ragan et al. 2015).

As mentioned above, we have 54 cores that are detected in at least one of the three deuterated species. For those detected in more than one line, we define the core velocity, vc, as an average of the detected centroid velocities, weighted by their measurement uncertainties.

The core velocities vc are illustrated in Figure 11. For cores with only one C18O(2–1) component, we compare the difference in centroid velocity between C18O and vc directly. If multiple CO velocities are found along the line of sight, we assume the component closest to vc is the one associated with the core, following the discussion in Kirk et al. (2007). This comparison is shown in Figure 13.

Figure 13.

Figure 13. Left: distribution of differences between the velocity of dense cores (determined with deuterated tracers) and the centroid velocity of C18O. Right: histogram of the distribution. The green line shows a Gaussian fit to this distribution.

Standard image High-resolution image

The median value of the velocity offset is 0.01 km s−1, with a standard deviation of about 0.3 km s−1. The majority of cores (71%) have core and envelope velocity offsets less than the sound speed of the ambient medium (0.27 km s−1 for 20 K temperature). This percentage is higher than that in NGC 1333, for which Walsh et al. (2007) found that half of their cores have differences greater than the sound speed, but similar to that seen in the Perseus cloud (Kirk et al. 2007). As discussed in Walsh et al. (2004), small relative motions between cores and envelopes could be interpreted as an indication of quiescence on small scales, and this would appear to argue against a competitive accretion scenario for star formation (Bonnell & Bate 2006), in which dense cores gain most of their mass by sweeping up material as they move through the cloud.

5.5. Virial State of Dense Cores

We now examine the dynamical state of the dense cores, i.e., the comparison of their internal kinetic energy (EK) and gravitational energy (EG). This ratio is captured by the virial parameter (Bertoldi & McKee 1992), defined as

Equation (4)

where σc is the intrinsic 1D velocity dispersion of the core and Rc is the core radius. The dimensionless parameter a accounts for modifications that apply in the case of nonhomogeneous and nonspherical density distributions. For a spherical core with a radial density profile that is a power law $\rho \propto {r}^{-{k}_{\rho }}$, then, for kρ = 0, 1, 1.5, 2, a = 1, 10/9, 5/4, 5/3. We adopt a fiducial value of kρ = 1.5 and a = 5/4, following McKee & Tan (2003). For a self-gravitating, unmagnetized core without rotation, a virial parameter above a critical value αcr = 2a indicates that the core is unbound and may expand, while one below αcr suggests that the core is bound and may collapse.

We measure core 1D velocity dispersions, σc, from each of the three dense gas tracers, i.e., ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$, ${\mathrm{DCO}}^{+}$, and $\mathrm{DCN}$. As shown above, their line widths can vary for the same core, so we calculate the virial parameters separately using each tracer. We derive the intrinsic velocity dispersion from the observed dispersion following Equation (3) (replacing C18O with other species). For the core masses, we use the values estimated assuming a temperature of 20 K, as listed in Appendix A.

For core radius, we attempt two methods. The first is to use the effective radius calculated from the dendrogram-returned area (see Section 5.1). For the second method, we adopt a deconvolved size defined as Rc = $\sqrt{(A-{A}_{\mathrm{beam}})/\pi }$, where A and Abeam are the core area and synthesized beam size, respectively. Note that in our core identification process, we have allowed for cores with an area smaller than the synthesized beam size. Here, for the virial analysis, we ignore the cores with areas smaller than 1.5 × Abeam, for which the deconvolved sizes could have very large uncertainties. This criterion excludes 34 out of 76 cores.

Figures 14(a) and (b) display the virial parameters measured with different tracers versus core mass for the two methods described above. In Figures 14(c) and (d), we combine the measurements from different tracers by taking the linear average of their nonthermal velocity dispersion in the virial parameter derivation.

Figure 14.

Figure 14. (a) Virial parameter vs. core mass, with radius measured from the dendrogram-defined area and velocity dispersion measured with different dense gas tracers, as shown in the legend. The critical value of ${\alpha }_{\mathrm{cr}}=2a\to 2.5$ is shown by the upper dashed line; cores below this line are gravitationally bound. The lower dashed line shows the virial equilibrium case of $\alpha =a\to 5/4$. (b) Same as (a) but with core radius estimated after allowing for beam deconvolution. Small cores, i.e., with areas >1.5Abeam, are excluded. (c) Same as (a), but we take the linear average of the nonthermal line width measured via different tracers to derive an average virial parameter. (d) Same as (c) but using the deconvolved size.

Standard image High-resolution image

We see virial parameters range from 0.5 to 10 as measured by individual dense gas tracers. There is a trend for more massive cores to have smaller virial parameters, but this is generally expected, since $\alpha \propto {M}_{c}^{-1}$. The scatter is significantly reduced for the deconvolved size method, with most measurements ranging from 0.5 to 3. This suggests that most data points with virial parameters >5 in panel (a) could arise from overestimation in the core radius. We do not find significant systematic differences between different tracers. The median values are 1.35, 1.19, and 1.23 for ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$, ${\mathrm{DCO}}^{+}$, and $\mathrm{DCN}$, respectively.

The virial parameters estimated by averaging all of the available dense gas data for each core show a further reduction in the scatter. For the second method with deconvolved sizes that focus on the larger cores, we obtain a median value of 1.22 and a standard deviation of 0.88. Most cores have a virial parameter that is consistent with a value expected in virial equilibrium, given the uncertainties.

The uncertainties in the derived virial parameters come from uncertainties in measured 1D line dispersion σobs, mass, and temperature. The fitting error of σobs is typically ≲20%, resulting in ∼40% uncertainty in σobs2. The assumed temperature will systematically affect the estimation of the dense core mass and also the thermal line width component in Equation (3). For example, with a typical ${\sigma }_{{\mathrm{DCO}}^{+}}$ = 0.36 km s−1, if temperatures of 15 or 30 K were to be adopted, then the virial parameters would differ by factors of 0.6 and 1.9, respectively. Also considering other uncertainties in the mass estimate, like dust opacity, gas-to-dust mass ratio, dust emission fluxes, and distances, overall, we estimate the absolute virial parameter uncertainties to be about a factor of 2.5. However, this uncertainty factor is itself quite uncertain and includes systematic effects, some of which are not expected to vary that much from core to core.

Given the best case derived average value of αvir = 1.22, we conclude that the dense cores in G286, which span a wide range of masses from ≲1 to ∼100 M, have properties that are consistent with them being close to virial equilibrium. In this case, self-gravity would have been important in forming the cores, and a basic assumption of the turbulent core accretion model of star formation (McKee & Tan 2003) would be confirmed. However, given the potentially large systematic uncertainties, it is difficult to be more certain about whether the dense cores are actually closer to a supervirial or subvirial state or whether magnetic fields are playing a role in supporting the cores. The situation could be improved in the future with accurate core-scale temperature and magnetic field measurements.

5.6. Core-to-core Velocity Dispersion

The relative motion between dense cores can be quantified using the core-to-core velocity dispersion σc−c, i.e., the standard deviation of the core centroid velocities. It can be compared with the velocity dispersion of the large-scale diffuse gas out of which these dense cores presumably formed or the initial velocity dispersion of newborn stars and, as such, provides important constraints on theoretical models and simulations of star cluster formation (e.g., Kirk et al. 2010; Foster et al. 2015).

Here our target G286 offers an interesting case of a massive protocluster that is still in the gas-dominated phase and actively forming stars. To measure the core velocity dispersion, we show the core velocity distributions measured with C18O(2–1), ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$(3–2), ${\mathrm{DCO}}^{+}$(3–2), and $\mathrm{DCN}$(3–2) in Figure 15. For comparison, the large-scale TP spectra of each line are also overlaid. The results combining velocities measured with deuterated tracers (54 cores; see Section 5.4) are also displayed in Figure 15(a). We then calculate the standard deviation of these distributions, obtaining 1.27 ± 0.11 km s−1 for the C18O-detected sample, 1.52 ± 0.21 km s−1 for the ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$ sample, 1.40 ± 0.15 km s−1 for the ${\mathrm{DCO}}^{+}$ cores, 1.50 ± 0.20 km s−1 for the $\mathrm{DCN}$ cores, and 1.39 ± 0.13 km s−1 for the combined results. The uncertainties here only account for sampling errors due to limited sample size, assuming the data points are drawn from a normal distribution.

Figure 15.

Figure 15. (a) Distribution of C18O(2–1) core centroid velocities (black). Overlaid is the TP C18O(2–1) spectrum (averaged over a 2farcm5 radius aperture) for comparison. The results of core velocities measured with deuterated species are shown by the magenta histogram. (b) Same as (a) but only for ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$(3–2). (c) Same as (a) but only for ${\mathrm{DCO}}^{+}$(3–2). (d) Same as (a) but only for $\mathrm{DCN}$(3–2). The TP $\mathrm{DCN}$(3–2) spectrum is averaged over a 1' radius aperture.

Standard image High-resolution image

In contrast to previous results in nearby cluster-forming clouds like Ophiuchus and Perseus (André et al. 2007; Kirk et al. 2007, 2010), our core velocities cover a wide range from −22.5 to −17 km s−1, and the distribution is not well approximated with a single Gaussian component. This is particularly clear for ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$ and ${\mathrm{DCO}}^{+}$; for these two tracers, the core velocities exhibit a bimodal distribution with two velocity groups, which agrees well with the averaged TP spectra. The $\mathrm{DCN}$ picks up core velocities in a relatively uniform pattern, filling in the gap around −19 km s−1; hence, the distribution combining all of the deuterated species is flatter, though more cores still cluster in the "blue" group at ∼−21 km s−1. On the other hand, though the C18O profile can be characterized with a Gaussian (with some skewness) peaking aroud −20 km s−1, and we do have more C18O-detected cores close to the systemic velocity (−20 to −19 km s−1), the C18O-measured core velocity distribution is still relatively flat. This indicates that the core-to-core velocity dispersion we measured here is largely contributed by the global velocity patterns.

The core velocity dispersion σc−c can be compared with the dispersion required for virial equilibrium on the protocluster clump scale, σcl,vir, and its actual gas velocity dispersion, σcl. For σcl,vir, we again follow Bertoldi & McKee (1992):

Equation (5)

As with cores, we again adopt kρ = 1.5, so that a = 5/4. We choose a size of Rcl = 1.54 pc, which is the geometric mean of the major and minor axes of the large ellipse boundary shown in Figure 1. From the Herschel SED-derived mass surface density map, we obtain a mass for this region of ∼2900 M. Thus, σcl,vir = 1.42 ± 0.36 km s−1, where the error comes assuming a 50% uncertainty in the mass estimate. We also tried a smaller aperture (by using major and minor axes that are a factor of 1/$\sqrt{2}$ smaller than the current ones) to more closely encompass the region containing dense cores, which gives σcl,vir = 1.54 ± 0.39 km s−1. The mass here only accounts for the gas component, since we do not expect a significant contribution from stellar mass: Andersen et al. (2017) estimated a total current stellar mass of ∼240 M in a similarly sized region. Thus, the observed values of σc−c are comparable or slightly smaller than σcl,vir, depending on which tracer is used.

For σcl, we measure the line width of the average TP spectra of C18O(2–1) in this region. The purpose here is to compare core-to-core motions with the spread of motions seen over the region as a whole to reveal how connected the dense cores are to the larger-scale gas in the region. A Gaussian fit of the C18O(2–1) line gives ${\sigma }_{\mathrm{cl},{{\rm{C}}}^{18}{\rm{O}}}$ = 1.09 ± 0.01 km s−1. To account for the thermal component, we correct this value following Equation (3) assuming a temperature of 20 K and obtain σcl = 1.12 km s−1.

In summary, the 1D dispersion measured in gas tracers, σcl (1.12 ± 0.01 km s−1), is slightly smaller than but still consistent with σcl,vir (1.48 ± 0.37 km s−1), indicating that the G286 clump could be in approximate virial equilibrium (assuming it is a single, coherent dynamical system) or modestly subvirial, but it is hard to be more precise given the uncertainties. We also see a range of σc−c values using different tracers, and the value using C18O(2–1), which includes most cores in this sample, i.e., ${\sigma }_{{\rm{c}}-{\rm{c}},{{\rm{C}}}^{18}{\rm{O}}}$ = 1.27 ± 0.11 km s−1, is similar to σcl,vir but slight larger than σcl. Here the dense core velocity distribution is more flat, while both the C18O core velocities and the TP C18O spectrum cover a similar velocity range. This means there is a deficiency of dense cores near the systemic velocity (∼20 km s−1), where the bulk of the C18O-traced gas is located. This deficiency is clearer in the distributions traced by ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$ and ${\mathrm{DCO}}^{+}$; hence, an even larger σc−c is measured with these two tracers. The two velocity groups seen in ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$ and ${\mathrm{DCO}}^{+}$ (at ∼−21 and −18 km s−1) are actually spatially distinct (see Figures 4, 8, and 11), with the more redshifted cores mostly located in the NE–SW filament and the more blueshifted cores in the NW–SE and E–W filaments. A similar velocity pattern is also seen for C18O in Figure 4, indicating that the dense cores are still well coupled with the large-scale motions within the cloud.

To better characterize the core-to-core velocity dispersions of the two velocity groups, we adopt a simple boundary to spatially differentiate them, shown as the dashed green line in the right panel of Figure 11. The blueshifted subsample (NW of the boundary) has a core-to-core velocity disperion σc−c,blue of 0.83 ± 0.10 km s−1 for C18O-detected cores (or 0.78 ± 0.10 km s−1 for measurements combining ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$, ${\mathrm{DCO}}^{+}$, and $\mathrm{DCN}$). Similarly, we have σc−c,red = 0.75 ± 0.10 km s−1 for C18O and 0.79 ± 0.12 km s−1 for the combined dense gas results. For the dispersion required for virial equilibrium, σcl,vir, we estimate the mass and radius using two approximate ellipse boundaries for each velocity group (shown in Figure 1), which yields σcl,vir,red = 0.96 ± 0.24 and σcl,vir,blue = 1.48 ± 0.37 km s−1. These numbers are also summarized in Table 2. Therefore, at least for the "blue" group, the core velocity dispersion σc−c is appears to be smaller than σcl,vir, potentially indicating that it is kinematically cold and subvirial, perhaps due to coherent motions within a filament. We also see that the dispersion in core velocities for the whole cloud mainly arises from the global velocity pattern of the red and blue groups.

Table 2.  Comparison of Core-to-core Velocity Dispersion and Velocity Dispersion Required for Virial Equilibrium (in km s−1)

  ${\sigma }_{{\rm{c}}-{\rm{c}},{{\rm{C}}}^{18}{\rm{O}}}$ σc−c,deua σvir
Blue group 0.83 ± 0.10 0.78 ± 0.10 1.48 ± 0.37
Red group 0.75 ± 0.10 0.79 ± 0.12 0.96 ± 0.24
Total 1.27 ± 0.11 1.39 ± 0.13 1.42 ± 0.36

Note.

aFor core velocity measurements combining deuterated species of ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$, ${\mathrm{DCO}}^{+}$, and $\mathrm{DCN}$.

Download table as:  ASCIITypeset image

The origin of this velocity pattern is uncertain. In the filament-collapse scenario, as observed in some hub–filament systems, accretion flows are channeling gas to the junctions where star formation is often most active (e.g., Kirk et al. 2013; Peretto et al. 2014; Liu et al. 2016). It is possible that these converging flows are reflected in different line-of-sight velocities depending on the 3D configurations. We will presumably have more massive cores in the hub region (near the systemic velocity) but not necessarily a larger number of cores, as suggested by our observations. Smoothly varying velocities along filaments are expected in this picture. We do see indications of a velocity gradient of dense cores along the filaments, but it is not clear in C18O, for which the spectra are often complex. Further higher-sensitivity observations of ${{\rm{N}}}_{2}{{\rm{H}}}^{+}$ and ${\mathrm{NH}}_{3}$ will help investigate the gas velocity gradient along filaments.

Alternatively, the two main velocity components seen in ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$ and ${\mathrm{DCO}}^{+}$ could be tracing two interacting clouds/filaments, with the central region as the collision interface (e.g., Nakamura et al. 2014). Such a mechanism could be consistent with a larger-scale cloud–cloud collision scenario that has been reported in other star-forming regions (e.g., Furukawa et al. 2009; Fukui et al. 2014; Gong et al. 2017).

Andersen et al. (2017) analyzed the stellar population in G286 and found evidence for at least three different subclusters associated with the molecular clump based on differences in extinction and disk fractions. It is unclear how the dense gas distribution and ongoing cluster formation might be related to these past star formation events. Future studies of the radial velocity of optically revealed stars, e.g., the velocity dispersion and its distribution, will be of great interest to understanding the cluster formation in G286.

6. Conclusion

We have studied the gas kinematics and dynamics of the massive protocluster G286 with ALMA using spectral lines of C18O(2–1), ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$(3–2), ${\mathrm{DCO}}^{+}$(3–2), and DCN(3–2). The main results are as follows.

  • 1.  
    Morphologically, C18O(2–1) traces more extended emission, while ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$(3–2) and ${\mathrm{DCO}}^{+}$(3–2) are more closely associated with the dust continuum. Here DCN(3–2) is strongly concentrated toward the protocluster center, where no or only weak detection is seen for ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$ and ${\mathrm{DCO}}^{+}$, possibly due to a relatively evolved evolutionary stage in the central region involving chemical evolution at higher temperatures.
  • 2.  
    Based on the 1.3 mm continuum, G286 is composed of several parsec-scale filamentary structures: the NE–SW filament in the NW and the NW–SE filament in the SE, as well as another filament oriented in the E–W direction that is more clearly seen in ${\mathrm{DCO}}^{+}$. The NE–SW filament is associated with redshifted C18O emission, while the NW–SE and E–W filaments are mainly associated with blueshifted gas. Other tracers show similar velocity structures.
  • 3.  
    We performed a filamentary virial analysis toward the NE–SW filament. We divided the filament into four strips, and the values of mf/mvir,f of the four strips range from 0.60 to 2.39. Within the uncertainties, these values are consistent with the filament being in virial equilibrium, without accounting for surface pressure and magnetic support terms. We also detected a steady velocity gradient of 2.84 km−1 pc−1 along the filament, which may arise from infall motion.
  • 4.  
    We analyzed the spectra of 74 continuum dense cores and measured their centroid velocities and internal velocity dispersions. There are no statistically significant velocity offsets among different tracers. Here C18O has a systematically larger velocity dispersion compared with other tracers.
  • 5.  
    The majority (71%) of the dense cores have subthermal velocity offsets with respect to their surrounding C18O-emitting envelope gas, similar to what was found in previous studies for low-mass star formation environments (e.g., Kirk et al. 2007).
  • 6.  
    We measured the virial parameters of the dense core in G286, which spans more than 2 orders of magnitude in mass. The average value of these virial parameters is close to unity, suggesting that the cores are close to virial equilibrium and self-gravity has been important for forming the cores. However, this conclusion is subject to revision if there is a large systematic error in the mass estimates of the cores.
  • 7.  
    The core-to-core velocity dispersion in G286 is similar to that required for virial equilibrium in the protocluster potential but with the velocity distribution largely composed of two spatially distinct velocity groups. This indicates that the dense molecular gas has not yet relaxed to virial equilibrium in the protocluster potential, even though the total velocity dispersion would indicate such a condition. The same analysis for the subregions corresponding to each velocity group reveals smaller core-to-core velocity dispersions, with one case consistent with virial equilibrium and the other case indicating a subvirial state.

This paper makes use of the following ALMA data: ADS/JAO.ALMA#2015.1.00357.S. ALMA is a partnership of the ESO (representing its member states), NSF (USA), and NINS (Japan), together with the NRC (Canada), NSC and ASIAA (Taiwan), and KASI (Republic of Korea), in cooperation with the Republic of Chile. The Joint ALMA Observatory is operated by the ESO, AUI/NRAO, and NAOJ. The National Radio Astronomy Observatory is a facility of the National Science Foundation operated under cooperative agreement by Associated Universities, Inc.

Appendix A: Properites of Dense Cores in G286

 We list in Table A1 the estimated physical parameters for the dense core sample of G286.

Table A1.  Estimated Physical Parameters for 1.3 mm Continuum Cores

Core R.A. Decl. Ipeak Sν Mc Rc Σc nH,c
  (deg) (deg) (mJy beam−1) (mJy) (M) (0.01 pc) (g cm−2) (106 cm−3)
1 159.63383 −58.31897 60.14 420.29 80.24 3.63 4.068 11.71
2 159.63524 −58.32063 15.32 47.19 9.01 1.74 1.994 11.99
3 159.64045 −58.32043 11.20 34.13 6.52 1.92 1.179 6.41
4 159.64373 −58.32201 11.10 34.49 6.58 1.76 1.416 8.39
5 159.63973 −58.32167 11.06 69.42 13.25 2.71 1.209 4.66
6 159.63154 −58.31674 10.03 14.94 2.85 1.05 1.713 16.97
7 159.64328 −58.32094 9.43 29.04 5.54 1.68 1.308 8.12
8 159.63153 −58.31933 9.04 7.07 1.35 0.73 1.694 24.24
9 159.64708 −58.32530 8.73 20.00 3.82 1.43 1.244 9.07
10 159.63546 −58.32133 8.52 5.40 1.03 0.66 1.569 24.74
11 159.63163 −58.31720 8.38 3.23 0.62 0.51 1.574 32.16
12 159.63177 −58.31842 8.30 5.69 1.09 0.68 1.570 24.11
13 159.63045 −58.31534 8.29 14.93 2.85 1.34 1.061 8.28
14 159.63572 −58.31830 7.92 5.31 1.01 0.70 1.380 20.58
15 159.66617 −58.32238 7.90 13.47 2.57 1.56 0.710 4.77
16 159.63329 −58.32012 7.57 7.69 1.47 0.83 1.429 18.01
17 159.62948 −58.31815 7.32 6.79 1.30 0.81 1.325 17.11
18 159.66145 −58.32416 7.31 9.14 1.74 1.26 0.740 6.16
19 159.63511 −58.31409 6.87 62.38 11.91 3.09 0.833 2.82
20 159.63008 −58.31798 6.67 6.77 1.29 0.83 1.258 15.86
21 159.63146 −58.31591 6.52 12.32 2.35 1.23 1.045 8.90
22 159.64112 −58.31903 6.45 21.01 4.01 1.79 0.835 4.87
23 159.62922 −58.31562 6.31 14.14 2.70 1.36 0.975 7.49
24 159.63281 −58.31702 6.16 3.02 0.58 0.58 1.144 20.61
25 159.64503 −58.32397 6.12 13.86 2.65 1.44 0.856 6.23
26 159.63331 −58.31758 6.08 3.27 0.62 0.60 1.160 20.21
27 159.63752 −58.31851 6.06 8.92 1.70 1.13 0.896 8.30
28 159.64744 −58.32461 5.79 5.82 1.11 0.87 0.987 11.89
29 159.64422 −58.32289 5.28 4.33 0.83 0.79 0.875 11.52
30 159.64468 −58.32329 5.22 3.61 0.69 0.73 0.866 12.39
31 159.62961 −58.31916 5.13 6.24 1.19 0.91 0.968 11.16
32 159.63175 −58.32042 5.12 5.04 0.96 0.89 0.814 9.57
33 159.62987 −58.32137 4.50 12.73 2.43 1.45 0.769 5.53
34 159.64877 −58.32960 4.43 7.77 1.48 1.34 0.555 4.34
35 159.63103 −58.32106 4.16 2.89 0.55 0.71 0.724 10.60
36 159.62937 −58.32062 4.12 2.70 0.52 0.68 0.745 11.45
37 159.63976 −58.32276 4.02 8.54 1.63 1.45 0.516 3.71
38 159.63000 −58.32016 4.00 5.38 1.03 0.96 0.737 7.99
39 159.63706 −58.31779 3.92 3.64 0.70 0.87 0.614 7.38
40 159.62838 −58.32134 3.86 2.22 0.42 0.62 0.731 12.26
41 159.63368 −58.31362 3.81 5.43 1.04 1.07 0.610 5.98
42 159.62900 −58.31655 3.77 6.43 1.23 1.16 0.612 5.52
43 159.64888 −58.32789 3.77 5.20 0.99 1.08 0.568 5.49
44 159.64251 −58.31957 3.69 6.50 1.24 1.12 0.660 6.15
45 159.63874 −58.31673 3.64 8.38 1.60 1.44 0.514 3.72
46 159.67328 −58.32603 3.63 5.34 1.02 1.17 0.501 4.49
47 159.63370 −58.31682 3.61 1.91 0.36 0.62 0.643 10.93
48 159.64524 −58.32276 3.46 3.46 0.66 0.90 0.549 6.40
49 159.64849 −58.32509 3.43 3.86 0.74 0.91 0.599 6.90
50 159.63991 −58.32611 3.42 4.53 0.86 1.11 0.472 4.45
51 159.67270 −58.32482 3.38 9.11 1.74 1.56 0.477 3.20
52 159.64614 −58.32435 3.32 1.72 0.33 0.60 0.608 10.60
53 159.64167 −58.31675 3.28 18.88 3.60 2.32 0.449 2.03
54 159.64989 −58.32882 3.27 4.23 0.81 1.00 0.545 5.72
55 159.64225 −58.32300 3.23 1.51 0.29 0.58 0.571 10.29
56 159.63973 −58.32419 3.18 3.90 0.74 0.98 0.517 5.50
57 159.63632 −58.32468 3.12 3.64 0.69 1.01 0.460 4.78
58 159.65086 −58.32812 2.98 3.46 0.66 0.99 0.454 4.82
59 159.63897 −58.32474 2.94 5.99 1.14 1.23 0.502 4.25
60 159.63765 −58.32215 2.90 2.61 0.50 0.88 0.432 5.14
61 159.63372 −58.31557 2.87 1.43 0.27 0.63 0.460 7.63
62 159.63206 −58.32324 2.84 1.69 0.32 0.68 0.467 7.17
63 159.64094 −58.30845 2.83 5.51 1.05 1.25 0.451 3.77
64 159.67355 −58.32445 2.70 1.74 0.33 0.72 0.425 6.13
65 159.64838 −58.31984 2.68 5.75 1.10 1.31 0.425 3.38
66 159.64144 −58.32606 2.67 4.48 0.86 1.17 0.419 3.75
67 159.63598 −58.31504 2.64 1.38 0.26 0.64 0.429 6.98
68 159.67361 −58.32257 2.61 1.94 0.37 0.76 0.431 5.93
69 159.64169 −58.32479 2.60 1.86 0.36 0.74 0.430 6.05
70 159.63645 −58.32321 2.56 6.47 1.23 1.39 0.424 3.18
71 159.63106 −58.32798 2.56 1.49 0.29 0.66 0.434 6.84
72 159.63961 −58.30961 2.55 1.86 0.36 0.75 0.427 5.99
73 159.63557 −58.31262 2.52 1.20 0.23 0.58 0.460 8.35
74 159.63958 −58.31643 2.49 1.31 0.25 0.62 0.436 7.35
75 159.64786 −58.32908 2.47 0.99 0.19 0.53 0.444 8.69
76 159.63002 −58.32627 2.39 1.16 0.22 0.59 0.423 7.46

Download table as:  ASCIITypeset images: 1 1

Appendix B: Spectral Fitting of the Core Sample

 We show in Figures B1 and B2 the extracted spectra of the dense core sample of G286 from different species. The estimated line parameters are also displayed.

Figure B1.

Figure B1. Spectra of C18O(2–1), ${{\rm{N}}}_{2}{{\rm{D}}}^{+}$(3–2), ${\mathrm{DCO}}^{+}$(3–2), and $\mathrm{DCN}$(3–2) of 76 continuum cores shown in different colors. The core masses are labeled at the top left. For the spectrum with a peak flux greater than 4σ, we perform a Gaussian fitting. The returned parameters (centroid velocity, velocity dispersion) for each line are displayed at the top left, in the same color as the corresponding line. The dashed vertical lines indicate the centroid velocity from line fitting. If there are multiple components for C18O, only the main component (the one closer to the other dense gas tracers; see text) is shown.

Standard image High-resolution image
Figure B2.

Figure B2. Continuation of Figure B1.

Standard image High-resolution image
Please wait… references are loading.
10.3847/1538-4357/ab879f