A publishing partnership

Identification of Methyl Isocyanate and Other Complex Organic Molecules in a Hot Molecular Core, G31.41+0.31

, , , , , , and

Published 2021 February 3 © 2021. The American Astronomical Society. All rights reserved.
, , Citation Prasanta Gorai et al 2021 ApJ 907 108 DOI 10.3847/1538-4357/abc9c4

Download Article PDF
DownloadArticle ePub

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

0004-637X/907/2/108

Abstract

G31.41+0.31 is a well known chemically rich hot molecular core (HMC). Using Band 3 observations from the Atacama Large Millimeter Array (ALMA), we have analyzed the chemical and physical properties of the source. We have identified methyl isocyanate (CH3NCO), a precursor of prebiotic molecules, toward the source. In addition to this, we have reported the presence of complex organic molecules (COMs) like methanol (CH3OH), methanethiol (CH3SH), and methyl formate (CH3OCHO). Additionally, we have used transitions from molecules like HCN, H13CO+, and SiO to trace the presence of infall and outflow signatures around the star-forming region. For the COMs, we have estimated the column densities and kinetic temperatures, assuming molecular excitation under local thermodynamic equilibrium (LTE) conditions. From the estimated kinetic temperatures of certain COMs, we found that multiple temperature components may be present in the HMC environment. Comparing the obtained molecular column densities between the existing observational results around other HMCs, it seems that the COMs are favorably produced in the hot core environment (∼100 K or higher). Though the spectral emissions toward G31.41+0.31 are not fully resolved, we find that CH3NCO and other COMs are possibly formed on grain/ice phase and populate the gas environment similar to other hot cores like Sgr B2, Orion KL, and G10.47+0.03.

Export citation and abstract BibTeX RIS

1. Introduction

Chemistry in high mass star-forming regions (HMSFRs) is very rich and it has a significant impact on the evolution of the interstellar medium (ISM; e.g., van Dishoeck & Blake 1998; Tan et al. 2014). Over the years, numerous complex organic molecules (COMs) have been identified in hot molecular cores (HMCs; e.g., Herbst & van Dishoeck 2009). Hot molecular cores are compact, hot, dense, and associated with luminous infrared (IR) or with ultra-compact (UC) H ii regions (Cesaroni et al. 2010; Bonfand et al. 2017). To date, a number of HMCs were discovered (Kurtz et al. 2000), while a limited number of circumstellar disks were found around B-type stars such as the disks in Cepheus A HW2 (Patel et al. 2005), HH80-81 (Girart et al. 2018), and the disk in the late-O-type star, IRAS 13481-6124 (Kraus et al. 2010). The majority of these disk structures have been identified by their dust emission. However, they are also associated with molecular emission. A complex organic molecule, CH3CN, was observed in Cepheus A HW2 and a P-V diagram of CH3CN transitions was used to understand the velocity gradient that explains the gravitationally bound rotational motion.

Many complex physical processes, such as accretion, infall, and outflows are present in the earliest phase of HMSFRs during their various evolutionary stages. Therefore the chemistry of these regions is used as a diagnostic tool to examine their nature. Complex organic molecules are omnipresent in HMCs and show rich chemistry. The chemistry of these regions plays an essential role in tracing the physical parameters associated with particular regions, such as density, temperature, and ionization rate. Since the abundances of chemical species are time dependent and the collapsing time of an HMSFR is short, a comparison of observed abundance with modeling can be used to trace the chemical diversity of those region. For HMCs, both grain surface and gas phase processes are efficient. Various desorption mechanisms, such as thermal (active in the hot core temperature) and non-thermal desorption (dominant at low temperature) processes help release the grain surface species to the gas phase.

G31.41+0.31 (hereafter G31) is an HMC that is thought to be located at a 7.9 kpc distance from the Earth (Churchwell et al. 1990). This source has a luminosity of about ∼3 × 105L. It is supposed to be heated by O- and B-type stars (Cesaroni et al. 2010; Beltrán et al. 2009). Previous studies (Cesaroni et al. 2011, 2017) found some hints of the existence of a disk in G31.41+0.31. Osorio et al. (2009) performed detailed modeling of the spectral energy distribution (SED) of dust to obtain the physical parameters of the star and the core. Using these parameters, they estimated the ammonia emission and compared it with the Very Large Array (VLA) observation having a sub-arcsec resolution of G31. They obtained the mass of the central star ∼20–25 M and the significantly higher mass (1400–1800 M) of the envelope by considering an envelope size of Renv = 30,000 au. Recently, Beltrán et al. (2018) observed G31 with the Atacama Large Millimeter/submillimeter Array (ALMA) at 1.4 mm and resolved G31 into two cores, namely the main core and a northeast (NE) core. They observed physical properties such as accelerating infall, rotational spin up, redshifted absorption, and possible outflow directions associated with this source. The UC H ii region is situated at the NE side of G31 and separated by ∼5'' from the main core (Cesaroni et al. 2010). The main core's estimated mass is 120 M, which is based on old distance of 7.9 kpc (Beltrán et al. 2018). However, a new parallax measurement suggests that G31 is located at a distance of 3.7 kpc. Therefore the luminosity of the source would be ∼4 × 104L and the estimated mass of the main core (Mc) of G31 is ∼26 M (Beltrán et al. 2019; Estalella et al. 2019; Reid et al. 2019). However, this estimation should be considered the lower limit because high angular resolution might have filtered out spatially extended emission (Beltrán et al. 2019).

A wide variety of species (around 70) have previously been identified in G31. The simplest form of sugar, glycolaldehyde (HOCH2CHO), was observed in G31 with the IRAM Plateau de Bure Interferometer (PdBI) at 1.4, 2.1, and 2.9 mm observations (Beltrán et al. 2009). This is the earliest evidence of the presence of sugar outside of our Galactic center. A large number of COMs such as methanol (CH3OH), methyl cyanide (CH3CN), ethanol (C2H5OH), ethyl cyanide (C2H5CN), methyl formate (CH3OCHO), dimethyl ether (CH3OCH3), ethylene glycol (CH2OH)2, and acetone (CH3COCH3) had already been identified in G31 by using both single dish (IRAM 30 m, GBT) and Submillimeter Array (SMA) observations (Isokoski et al. 2013; Rivilla et al. 2017, and references therein).

The presence of a large number of COMs, including branched chain molecules, might be an indication of the existence of prebiotic molecules and essential building blocks of life in space (Belloche et al. 2014; Chakrabarti et al. 2015; Majumdar et al. 2015; Sil et al. 2018; Sahu et al. 2020). Goesmann et al. (2015) detected various COMs (e.g., methyl isocyanate, formamide, glycolaldehyde, ethylamine) in a Jupiter family comet, 67P/Churyumov–Gerasimenko (67P/C–G). Among them, methyl isocyanate (CH3NCO) was found to be relatively abundant as compared to other observed COMs. However, Altwegg et al. (2017) reported the resulting CH3NCO abundance as an upper limit. The presence of CH3NCO in space is supposed to be important because of its peptide-like bond. Detection of CH3NCO was first claimed by Halfen et al. (2015) in a massive hot molecular core, Sgr B2. It was also detected in Orion KL (Cernicharo et al. 2016). It has recently been observed in G10.47+0.03 (Gorai et al. 2020). Here, in G31, we have identified some transitions of this species.

Methanethiol (CH3SH) is a sulfur analog of methanol, where the S atom is replaced by an O atom. To date, this is the only complex sulfur-bearing species (having six atoms) that has been firmly detected in the ISM. Tentative identification of higher-order thiols, such as ethanethiol (C2H5SH), has been reported in Sgr B2N (Müller et al. 2016). Identification of CH3SH was also reported toward the Galactic center (Linke et al. 1979; Müller et al. 2016), Orion KL (Kolesniková et al. 2014), the chemically rich massive hot core, G327.3-6 (Gibb et al. 2000), and 67P/C–G comet (Calmonte et al. 2016).

In this work, we present an interferometric (ALMA archival data) observation toward G31. We report the first detection of CH3NCO and CH3SH in G31 and discuss the kinematics associated with this HMC. We also discuss the chemistry of other complex species in this source. The remainder of this paper is summarized as follows. In Section 2, we describe observational details and data reduction processes. All results are presented and discussed in Sections 3 and 4, respectively. Finally, we draw our conclusion in Section 5.

2. ALMA Archival Data

In this paper, we used ALMA cycle 3 archival data (#2015.1.01193.S) to study COMs in G31. The observation was performed during June 2016 using 40 12 m antennas. The shortest and the longest baseline were 15 m and 783.0 m, respectively, and the synthesized beam size was ∼1farcs1. Four spectral windows (86.559–87.028 GHz, 88.479–88.947 GHz, 98.977–99.211 GHz, and 101.076–101.545 GHz) of width 0.47 GHz were set up to observe the target source, G31; the phase center was α(J2000) = 18h47m34fs309, δ = −1°12'46farcs00. The flux calibrator was Titan, the phase calibrator was J1824+0119, and the bandpass calibrator was J1751+0939. The data cube has a spectral resolution of 244 kHz (∼0.84 km s−1) with a synthesized beam of ∼1farcs19 × 0farcs98, having a position angle (PA) of ∼ 76°. The systematic velocity of the source is known to be ∼97 km s−1 (Cesaroni et al. 2010; Rivilla et al. 2017). Image analysis was done using CASA 4.7.2 software (McMullin et al. 2007). Implementing the first-order baseline fit and by using the "uvcontsub" command available in the CASA program, we separated each spectral window into two data: continuum and line emission. The identification of lines of all the observed species presented in this paper is carried out using CASSIS (this software was developed by IRAP-UPS/CNRS, http://cassis.irap.omp.eu) together with the spectroscopic database "Cologne Database for Molecular Spectroscopy" (CDMS; Müller et al. 2001, 2005)6 and the Jet Propulsion Laboratory (JPL; Pickett et al. 1998)7 database.

3. Results

3.1. Continuum Emission and H2 Column Density Estimation

The continuum image of G31 at 94.3 GHz is shown in Figure 1. The synthesized beam size of the continuum is 1farcs19 × 0farcs98 with a PA of ∼ 76°. Using a two-dimensional Gaussian fitting over the dust continuum emission, we obtained the deconvolved size of G31 ∼ 0farcs90 × 0farcs68.

Figure 1.

Figure 1. Continuum map of G31 obtained with ALMA at 94 GHz. Contour levels start at 4σ (1σ = 7 mJy/beam) and steps are in 3σ. Ellipse at the below left corner of the image shows the synthesized beam (1farcs19 × 0farcs98) with PA = 76fdg83.

Standard image High-resolution image

Assuming a distance to the source of 7.9 kpc, we obtained a source size of ∼7110 × 5372 au. This source size changes to 3330 × 2516 au when a distance of ∼3.7 kpc is used.

Beltrán et al. (2018) observed G31 with an angular resolution of 0farcs22 and found a continuum brightness temperature of ∼132 K, which is comparable to the dust temperature (∼150 K). This suggests that the dust continuum emission observed at 217 GHz is optically thick. Rivilla et al. (2017) also observed that dust opacity is high (τ = 2.6) at 220 GHz.

We have observed that the peak intensity of the continuum is ∼158 mJy/beam which corresponds to a brightness temperature of 18 K (using Rayleigh–Jeans approximation, 1 Jy/beam ≡ 118 K). This estimation is much smaller than the assumed dust temperature of 150 K, which indicates a small optical depth (τν) of ∼ 0.12 (using Tmb = Td(1−exp(−τν)).

It should be noted that in our observation, the source was not resolved or at best marginally resolved; therefore, the estimated brightness temperature could be underestimated. After adding the beam dilution correction for the unresolved source, the brightness temperature is estimated to be ∼36 K. Considering this value of brightness temperature, dust emission remains optically thin (τν ∼ 0.24 ). Due to the collapsing envelope, we can expect a higher density and temperature toward the inner region than the outer envelope of this source. Since dust density is higher toward the inner region, opacity would be higher, as observed by Beltrán et al. (2018). It may reduce the true line intensity of the observed transitions. However, due to the lower angular resolution, our present data would not provide further details about the dust opacity of the whole region of G31.

Assuming the dust continuum to be optically thin, flux density can be written as,

Equation (1)

where Ωbeam = $\tfrac{\pi }{4{ln}2}$ × θmajor × θminor is the solid angle of the synthesized beam, S${}_{\nu }^{\mathrm{beam}}$ is the peak flux density measured in units of mJy/beam, τν is optical depth, Td is the dust temperature, and Bν(Td) is the Planck function (Whittet 1992). Optical depth can be expressed as

Equation (2)

where ρd is the mass density of dust, κν is the mass absorption coefficient, and L is the path length. Using the dust-to-gas-mass ratio (Z), the mass density of the dust can be written as

Equation (3)

where ${\rho }_{{{\rm{H}}}_{2}}$ is the mass density of the hydrogen molecule, ${N}_{{{\rm{H}}}_{2}}$ is the column density of hydrogen, mH is the hydrogen mass, and μH is the mean atomic mass per hydrogen atom. Here, we used Z = 0.01, μH = 1.41 (Cox 2000), and a dust temperature of 150 K. A two-dimensional Gaussian fitting of the continuum yields a peak intensity (S${}_{\nu }^{\mathrm{beam}}$) of dust continuum of 158.4 mJy/beam and an integrated flux (Fν) of 242.4 mJy. The rms noise of the continuum map was found to be 5 mJy/beam. From the above equations, the column density of molecular hydrogen can be written as

Equation (4)

According to the interpolation of the data presented in Ossenkopf & Henning (1994), we estimated the mass absorption coefficient. By considering a thin ice condition here, we obtained the mass absorption coefficient at 94 GHz (3187.90 μm) ∼0.176 cm2 g−1. By using Equation (4), the total hydrogen column density of the G31 source is ∼1.53 × 1025 cm−2.

3.2. Line Analysis, Molecular Emission, and Spectral Fitting

Observed spectra were obtained within the circular region having a diameter of 1farcs1 centered at R.A. (J2000) = 18h47m34fs309, decl. (J2000) = −1°12'46farcs00. Line parameters of all the observed transitions were obtained using a single Gaussian fit. From fitting, we have estimated the line width (FWHM), LSR velocity, and the integrated intensity. All the line parameters of observed molecules such as molecular transitions (quantum numbers) along with their rest frequency (ν), line width (FWHM, ΔV), integrated intensity (∫TmbdV), upper state energy (Eu), VLSR, and transition line intensity (2, where S is the transition line strength and μ is the dipole moment) are presented in Table 1. Overall, we have observed a broad line width (≥5 km s−1) and all lines are at the center around the systematic velocity of the source (∼97 km s−1; see Table 1). In the Appendix, we have presented observed and Gaussian-fitted spectra (see Figures A2A6).

Table 1.  Summary of the Line Parameters of Observed Molecules toward G31.41+0.31

Species ($J{{\prime} }_{{K}_{{\rm{a}}}^{{\prime} }{K}_{{\rm{c}}}^{{\prime} }}$-$J{{\prime\prime} }_{{K}_{{\rm{a}}}^{{\prime\prime} }{K}_{{\rm{c}}}^{{\prime\prime} }}$) Frequency (GHz) Eu/kB (K) FWHM (km s−1) Peak Intensity, Iν (K) VLSR (km s−1) 2(Debye2) Tmbdv (K km s−1)
SO2 83,5 − 92,8 86.63908 55.20 8.46 14.63 97.27 3.01 131.9 ± 3.6
H2CO 61,5 − 61,6 101.33299 87.57 10.20 75.90 97.26 5.05 824.4 ± 22.4
                 
CH3SH 40,4 − 30,3 (A+) 101.13915a 12.14 8.70 6.00 97.00 6.61 55.2 ± 17.7
  40,4 − 30,3 (E+) 101.13965a 13.56 6.10 5.90 95.45b 6.61 38.5 ± 14.8
                 
  42,3 − 32,2 (A−) 101.15933c 31.26 6.75 3.50 97.00 4.96 26.1 ± 4.7
  43,2 − 33,1 (E−) 101.15999c 52.39 7.09 1.56 95.03d 2.90 13.1 ± 2.9
  43,1 − 33,0 (A+) 101.16066c 52.55 7.09 1.56 93.05d 2.90 13.1 ± 2.9
  43,2 − 33,1 (A−) 101.16069c 52.55 7.09 1.56 92.94d 2.90 13.1 ± 2.9
                 
  42,3 − 32,2 (E−) 101.16715e 29.62 6.94 8.20 97.00 4.96 60.6 ± 5.2
  42,2 − 32,1 (E+) 101.16830e 30.27 6.20 7.70 93.58f 4.96 50.8 ± 4.5
                 
CH3OH ${7}_{2,{5}^{-}}-{6}_{3,{3}^{-}},{{\rm{v}}}_{{\rm{t}}}=0$ 86.61557 102.70 9.06 84.55 97.02 1.35 815.3 ± 26.8
  ${7}_{2,{5}^{+}}-{6}_{3,{4}^{+}},{{\rm{v}}}_{{\rm{t}}}=0$ 86.90291 102.72 9.13 85.16 97.00 1.35 827.8 ± 24.9
  ${15}_{3,{13}^{+}}-{14}_{4,{10}^{+}},{{\rm{v}}}_{{\rm{t}}}=0$ 88.59478 328.26 8.98 82.26 97.10 4.20 786.4 ± 14.6
  ${15}_{3,{12}^{+}}-{14}_{4,{11}^{-}},{{\rm{v}}}_{{\rm{t}}}=0$ 88.93997 328.28 8.55 79.78 97.16 4.20 726.5 ± 9.9
  5−2,3 − 51,4, E2, vt = 0 101.12685 60.73 8.37 23.87 98.88 0.01 208.4 ± 18.8
  7−2,5 − 71,6, E2, vt = 0 101.29341 90.91 7.72 45.71 97.38 0.05 373.2 ± 18.0
  8−2,6 − 81,7, E2, vt = 0 101.46980 109.49 7.25 55.31 97.98 0.09 426.9 ± 16.5
                 
CH3NCO 100,0 − 90,0 86.68655 34.97 7.70 14.63 97.83 83.11 120.0 ± 6.8
  100,10 − 90,9 86.68019 22.88 8.28 13.50 97.90 83.06 119.1 ± 11.5
  102,9 − 92,8 86.78077 46.73 6.80 10.45 97.32 79.73 75.8 ± 4.5
  102,8 − 92,7 86.80502 46.74 5.82 11.16 97.10 79.74 69.2 ± 3.8
  10−3,0 − 9−3,0 86.86674 88.58 10.14 15.80 97.00 75.32 170.6 ± 16.6
  102,0 − 92,0 87.01607 58.88 6.22 8.91 97.12 79.18 59.0 ± 6.2
  100,10 − 90,9 86.67492 285.03 83.07
                 
CH3OCHO 173,14 − 173,15 86.91761 99.72 6.41 7.09 98.03 1.87 48.5 ± 7.6
  113,9 − 112,10 E 88.68688 44.97 5.41 12.90 97.86 2.44 74.3 ± 9.2
  113,9 − 112,10 A 88.72326 44.96 7.84 17.11 98.32 2.43 142.8 ± 29.0
  71,6 − 61,5 E 88.84318 17.96 8.07 45.87 98.18 18.04 394.3 ± 58.9
  71,6 − 61,5 A 88.85160 17.94 6.64 49.17 97.50 18.04 348.0 ± 22.4
  81,8 − 71,7 E 88.86241 207.10 5.83 23.49 97.64 20.89 145.9 ± 10.1
  133,11 − 132,12 E 101.37050 59.64 5.16 15.90 97.93 2.61 87.4 ± 6.7
  133,11 − 132,12 A 101.41474 59.63 5.93 14.56 97.81 2.60 92.0 ± 13.5
                 
SiO (2-1) 86.84696 6.25 94.1 2.82

Notes.

aSpectral profile 1. bVelocity with respect to 101.13915 GHz transition. cSpectral profile 2. dVelocity w.r.t 101.15933 GHz transition. eSpectral profile 3. fVelocity w.r.t 101.16715 GHz transition.

Download table as:  ASCIITypeset image

3.2.1. Simple Molecules

We have observed SO2 (86.63908 GHz) and H2CO (101.33299 GHz) in G31. We have also identified some simple molecules such as HCN, H13CO+, and SiO, which are widely used to trace various physical processes of star-forming regions. The observed spectral profile of H13CO+ shows redshifted absorption, which is shown in Figure 2(a). This is a signature of an inverse P-Cygni profile. This suggests that the source has an infall nature. Observed emission of SiO and HCN traces the outflows in this source.

Figure 2.

Figure 2. Observed spectra of (a) H13CO+ (inverse P-Cygni) and (b) HCN toward the dust continuum peak. 86.631847 GHz transition of HCN is used to set the velocity scale. The other two lines (green dashed) are plotted in the same figure by taking the difference in frequency between the transitions which corresponds to have a velocity shift of ∼7 km s−1 for transition 88.63393 and ∼5 km s−1 for 88.63041 GHz. The bold, dashed, green line shows the systematic velocity (VLSR) of the source ∼97 km s−1.

Standard image High-resolution image

Figure 2(b) shows the observed spectral profile of HCN. Here, we have identified three hyperfine transitions of HCN (F = 1 → 1, 2 → 1, and 1 → 0 hyperfine components of the J = 1 →  0 transition, see Table 2). Three hyperfine transitions of HCN appear in the absorption and are blended together. These three absorption lines together result in a broad feature. HCN traces the outflow associated with the source. The observed transitions of both HCN and H13CO+ are found to be optically thick.

Table 2.  Observed Molecular Transitions in Absorption

Species ($J{{\prime} }_{{K}_{{\rm{a}}}^{{\prime} }{K}_{{\rm{c}}}^{{\prime} }}$-$J{{\prime\prime} }_{{K}_{{\rm{a}}}^{{\prime\prime} }{K}_{{\rm{c}}}^{{\prime\prime} }}$) Frequency in GHz Eu (K)
H13CO+ 1-0 86.75430 4.16
HCN 1-0(F = 1-1) 88.63041 4.25
HCN 1-0(F = 2-1) 88.63184 4.25
HCN 1-0(F = 0-1) 88.63393 4.25

Download table as:  ASCIITypeset image

Here, we have also detected one transition of SiO at 86.84696 GHz (J = 2 → 1). SiO shows the absorption feature toward the center of the dust continuum and the emission profile at all other positions. Beltrán et al. (2018) also observed a similar spectral feature of SiO (5-4, 217.10498 GHz). SiO traces outflows associated with this source.

3.2.2. Complex Organic Molecules

We have observed numerous COMs that were previously reported in G31 (e.g., Rivilla et al. 2017, and references therein). Additionally, for the first time in this source, we are reporting the detection of two new molecules, methyl isocyanate (CH3NCO) and methanethiol (CH3SH). Altogether, we have identified six transitions of CH3NCO with the ground vibrational state (vb = 0) and one transition with vb = 1. Among the six transitions of CH3NCO, one transition (86.86674 GHz) is blended with ethylene glycol (confirmed from LTE modeling).

We have identified three different peaks of CH3SH, where each peak contains multiple transitions. These transitions are reported in Table 1. We are unable to separate these observed transitions. Because of closely spaced transitions, it is difficult to distinguish two or more different transitions from a single observed peak of CH3SH. Thus, to obtain the observed line parameters of CH3SH transitions, we have fitted multiple Gaussian components to the observed spectra. For multiple Gaussian fittings, we have used fixed values of velocity separation and the expected line intensity ratio. Only the FWHM is kept as a free parameter. This method works well in the observed spectral profiles 1 and 3 around 101.139 GHz and 101.167 GHz of CH3SH, where there are two transitions around each profile (see Table 1). Two-component Gaussian fitting corresponding to the two transitions is used to fit those spectral profiles (see Table 1 and Figure A5). However, this method does not work well for spectral profile 2, around 101.159 GHz, where four transitions are present. The four transitions can be classified into two classes based on their expected peak intensity (equivalent to 2) and Eup. The expected intensity ratio between these four transitions is 1:0.58:0.58:0.58. We have used two-component Gaussian fitting corresponding to the two classes of transitions, one for the 101.15933 GHz (2 = 4.96, Eup = 31.26 K) transition and another component for the other three transitions, 101.15999 GHz, 101.16066 GHz, 101.16069 GHz (2 = 2.90, Eup = 52.55 K) around the 101.159 GHz line profile. For the fitting, we have used a 1:1.74 (1.74 is the sum of the intensity of three transitions) intensity ratio for these two components. Therefore, the resultant Gaussian corresponding to the three transitions can be obtained via this method; as these transitions are equally intense and have a similar excitation temperature, the area under the Gaussian can be divided by three to get the contribution from each transition multiplet. We have used only one transition from these multiplets for the rotation diagram of CH3SH since the integrated area of all three transitions is the same and has similar upper state energy (Eup). Therefore, these fitting results are used meaningfully to estimate rotational temperature using the rotation diagram method. Obtained line parameters are further used in the rotation diagram of CH3SH. For the rotation diagram of CH3SH, we use 101.13915, 101.13965, 101.15933, 101.1599, 101.16715, and 101.16830 GHz transitions. Gaussian fits of CH3SH observed spectra are provided in Figure A5.

We also have identified the presence of two abundant COMs, methanol and methyl formate, in this source. Methanol is one of the most abundant COMs in star-forming regions. Here, we have identified seven transitions of CH3OH with different upper state energy in the ALMA band 3 regions. We have also identified CH3OCHO transitions with a wide range of upper-level energies (17–207 K), which are listed in Table 1. The integrated intensity maps of some unblended transitions of SO2, H2CO, CH3SH, CH3OH, CH3OHCO, and CH3NCO are shown in Figures 3(a)–(f) where the continuum is overlaid on the integrated intensity map, and all other maps corresponding to different frequencies are given in the Appendix (see Figures A9A12). Using local thermodynamic equilibrium modeling and comparing that with the integrated intensities of observed molecular transitions, we found that various transitions of CH3OH, CH3NCO, and CH3OCHO are optically thin.

Figure 3.

Figure 3. Integrated intensity distribution (color) of (a) SO2, (b) H2CO, (c) CH3OH, (d) CH3OCHO, (e) CH3SH, (f) and CH3NCO lines overlaid on the 3.1 mm continuum emission (black contours). Contour levels are at 20%, 40%, 60%, and 80% of the peak flux. The synthesized beam is shown in the lower left-hand corner of each figure.

Standard image High-resolution image

For example, in the LTE and optically thin condition, the expected line ratio between the 87.01608 GHz and 86.68019 GHz transitions of CH3NCO is 1.6 while the observed intensity (brightness temperature) ratio is 1.44. This suggests that these two lines are optically thin. Similarly, for 86.78078 GHz and 86.68019 GHz transitions of CH3NCO, the observed intensity ratio between these two lines is 1.35, and the expected line ratio is 1.30. For the 101.12685 GHz and 101.46980 GHz transitions of CH3OH, the observed intensity ratio is 2.3, which is consistent with the expected LTE line ratio of 2.2. For CH3OCHO, the observed line ratio between 88.68688 GHz and 88.85160 GHz transition is 4, which is similar to the observed integrated intensity ratio 3.7. The similarity between the expected and observed line ratios of these transitions suggest that they are optically thin.

3.3. Rotation Diagram Analysis: Estimation of Gas Temperature

We have detected multiple transitions of CH3OH, CH3OCHO, CH3SH, and CH3NCO, with different Eu. Thus, we employed the rotation diagram analysis to obtain the rotational temperatures. We performed the rotation diagram analysis by assuming the observed transitions are optically thin and in LTE. For optically thin lines, column density can be expressed as (Goldsmith & Langer 1999),

Equation (5)

where gu is the degeneracy of the upper state, kB is the Boltzmann constant, ∫TmbdV is the integrated intensity, ν is the rest frequency, μ is the electric dipole moment, and S is the transition line strength. Under LTE conditions, the total column density can be written as,

Equation (6)

where Trot is the rotational temperature, Eu is the upper state energy, Q(Trot) is the partition function at rotational temperature. The Equation (6) can be rearranged as,

Equation (7)

Equation (7) shows that there is a linear relationship between Eu and $\mathrm{log}({N}_{{\rm{u}}}/{g}_{{\rm{u}}})$. Two parameters, column density and rotational temperature, can be obtained by fitting a straight line to the values of $\mathrm{log}({N}_{{\rm{u}}}/{g}_{{\rm{u}}})$ plotted as a function of Eu, where Nu/gu is obtained from the observations through Equation (5).

Rotational diagrams of methyl isocyanate, methanol, methanethiol, and methyl formate are presented in Figure 4. For CH3OCHO, we have several transitions, but we do not see two components. Hence we fit all the data points with a line, which are not aligned. Though we did not find a very good fitting, the obtained rotational temperature of methyl formate ∼162 K is consistent with the previously measured value in this source for the same molecule (Beltrán et al. 2005; Rivilla et al. 2017). We have obtained a rotational temperature of 76 K for CH3SH and 48 K for CH3NCO. The rotational diagram analysis suggests two temperature components of methanol; one is a cold component and the other is a hot component. The cold component (solid black line, see Figure 4(b)) temperature is obtained 33 K. In contrast, the hot component (solid blue line, see Figure 4(b)) temperature is 181 K. The rotational diagram of CH3OH suggests that the transitions associated with the higher Eu transitions (86.61557 GHz, 86.90291 GHz, 88.59478 GHz, 88.93997 GHz; hot component) are mainly arising from the hot gas surrounding the G31. In contrast, the emissions associated with the lower Eu (101.12685 GHz, 101.29341 GHz,101.46980 GHz; cold component) transitions are from the cold region of the HMC. However, due to our observation's lower angular and spatial resolution, we marginally resolved these transitions. Therefore, it is not possible to provide a clear insight into the spatial distribution of methanol in this source.

Figure 4.

Figure 4. Rotational diagram of (a) CH3NCO, (b) CH3OH, (c) CH3SH, and (d) CH3OCHO. Black filled squares show the data points and the solid black and blue lines represent the fitted curve.

Standard image High-resolution image

The error bars (vertical red bars) in Figure 4 are the absolute uncertainty in a log of (Nu/gu), and this arises from the error of the observed integrated intensity that we measured using a single Gaussian fitting to the observed profile of each transition. The integrated intensities with their uncertainties in all observed transitions are provided in Table 1. Our estimated rotational temperatures for methyl formate and the hot component of methanol is similar to the typical hot core temperature. On the other hand, methyl isocyanate and the cold component of methanol have lower rotational temperatures (33–48 K). However, due to poor angular resolution, we could not provide further details about these two temperature components' spatial distribution.

To obtain the molecular column density of the observed chemical species, we used LTE modeling with CASSIS. For this modeling, we used the estimated rotational temperature discussed above. We calculated the beam average column density (Nb), where the effect of beam dilution was not taken into consideration. For the model, we used five parameters: (i) column density, (ii) excitation temperature (Tex), (iii) line width (ΔV), (iv) velocity (VLSR), and (v) source size (θs). To obtain the best fit with the observation, only column density was varied by keeping all other parameters (Tex, source size, FWHM, and hydrogen column density) fixed and we chose the best-fit results when LTE model spectra are the same as the observed spectra. For LTE fitting, we used source sizes for molecular emission similar to the deconvolved size we obtained from a 2D Gaussian fitting. Line width (FWHM) for molecular transitions is estimated from the Gaussian fitting of spectral profiles, and excitation temperature (Tex) is assumed to be ∼150 K, a typical hot core temperature. Fixed parameters are noted in Table 3. We did not have a rotation diagram for all the species (e.g., SO2, H2CO) presented here. Thus, we followed the LTE model to derive the column density of all the species. We have also compared the column densities of CH3OH, CH3SH, CH3NCO, and CH3OCHO between the rotation diagram analysis and the LTE model (see Table 4 and Section 3.6). Observed column densities and fractional abundances of all the species are presented in Table 4. Abundances of all species are noted with regard to H2.

Table 3.  Input Parameters of LTE Model

Species Source Size ('') FWHM (km s−1)
SO2 0.7 8.4
H2CO 1.3 10.2
CH3SH 0.8 5.5
CH3OH 1.3 8.5
CH3NCO 0.8 6.5
CH3OCHO 1.35 6.0

Note. H2 column density = 1.53 × 1025 cm−2 Tex = 150 K.

Download table as:  ASCIITypeset image

Table 4.  Comparison of Observed Molecular Column Density, Fractional Abundances, and Molecular Ratios between G31 and Other Sources

Species G31.41+0.31, Sgr B2 (N) Orion KL G10.47+0.03 Chemical Model
  Column Density Column Density Column Density Column Density Column Density
  [Abundance] [Abundance] [Abundance] [Abundance] [Abundance]
  LTE Rotation Diagram        
CH3OH 1.84(19)[1.2(−6)]a 2.94(19)[1.92(−6)] 4.00(19)[1.14(−5)]b 3.24(18)[1.62(−6]c 9.0(18)[1.8(−6)]d 7.56(19)[2.10(−5)]e
CH3OCHO 6.50(18)[4.25(−7)] 4.30(18)[2.81(−7)] 1.20(18)[3.33(−7)]f 1.04(17)[5.2(−8)]g 7.0(17)[1.4(−7)]d 3.31(17)[9.20(−8)]e
CH3NCO 7.22(17)[4.72(−8)]a 1.58(16)[1.04(−9)] 2.20(17)[6.11(−8)]h 7.0(15)[3.5(−9)]i 1.2(17)[8.88(−9)]j 4.68(16)[1.30(−8)]f
CH3SH 4.13(17)[2.70(−8)] 2.85(16)[1.86(−9)] 3.40(17)[9.44(−8)]b 2.0(15)[1.0(−9)]k 1.33(16)[3.70(−9)]b
SO2 2.88(18)[1.88(−7)]   8.28(16)[2.30(−8)]l 1.0(17)[5.0(−8)]m 3.0(17)[6.0(−8)]d  
H2CO 2.90(18)[1.90(−7)]   1.82(17)[5.06(−8)]n 1.5(17)[7.5(−8)]o 3.0(18)[6.0(−7)]d  
SiO 1.46(15)[9.54(−11)]p   4.32(15)[1.20(−9)]l 5.4(14)[2.7(−10)]q  
CH3SH/CH3OH 0.0225 0.0010 0.0082 0.0006   0.0002
CH3OCHO/CH3OH 0.3541 0.1463 0.0292 0.0321 0.0778 0.0044
CH3NCO/CH3OH 0.0393 0.0005 0.0054 0.0022 0.0049 0.0006

Notes. Orion KL (${N}_{{{\rm{H}}}_{2}}$) = 2.0 × 1024 cm−2, Sgr B2(N) (${N}_{{{\rm{H}}}_{2}}$) = 3.6 × 1024 cm−2, G10.47+0.03 (${N}_{{{\rm{H}}}_{2}}$) = 5.0 × 1024 cm−2, G31.41+0.31 (${N}_{{{\rm{H}}}_{2}}$) = 1.53 × 1025 cm−2.

aFor the hot component of methanol, the column density is derived using a higher value of excitation temperature T = 300 K. Using LTE, we have found opacities ≥1 for a few transitions of CH3OH (86.68655, 86.68019, 86.78077, and 86.80502 GHz) and CH3NCO (86.68019 and 86.68655 GHz). Thus, the column density estimated for CH3OH and CH3NCO by using LTE should be considered a lower limit. bMüller et al. (2016). cFavre et al. (2015). dRolffs et al. (2011). eGarrod (2013). fBelloche et al. (2016). gFavre et al. (2011). hBelloche et al. (2017). iCernicharo et al. (2016). jGorai et al. (2020). kKolesniková et al. (2014). lHiguchi et al. (2015). mTercero et al. (2018). nBelloche et al. (2013). oEsplugues et al. (2013). pSiO column density and fractional abundance calculated in the outflowing gas measured at VLSR = 94.1 km s−1 by averaging over the box (7'' × 7'') centered at (α(J2000), δ (J2000)) = [18h47m34fs121, −1°12'48farcs975]. qSchilke et al. (2001).

Download table as:  ASCIITypeset image

3.4. Radiative Transfer Modeling

Absorption studies are best suited to trace the physical conditions (e.g., density, kinetic temperature) of the ISM because it is not induced by beam dilution as long as the absorbing layer is larger than the continuum source. Here, we used non-LTE modeling to derive excitation conditions and physical properties from these absorption features. For the non-LTE modeling, we use the RADEX program (Van der Tak et al. 2007). The collisional rates were taken from the LAMDA database and the spectroscopic constants were taken from the JPL/CDMS catalog. In RADEX, we used a continuum background temperature (Tc) of 36 K, which corresponds to the peak brightness temperature of the 3.2 mm continuum, considering the beam dilution. For this modeling, we have used a colum density of 3.55 × 1015 cm−2, 8.92 × 1015 cm−2 and 1.18 × 1015 × cm−2 for H13CO+, HCN, and SiO respectively. Figures 5(a)–(c) show the variation of radiation temperature of HCN (1-0), H13CO+ (1-0), and SiO (2-1), respectively, for a wide range of parameter space. Our parameter space consists of the number density of hydrogen (varies in between 103 and 107 cm−3) and kinetic temperature (varies between 1 and 100 K). Figures 5(a)–(c) show the parameter space for the absorption features of HCN, H13CO+, and SiO, respectively. We have estimated the critical density of HCN (1-0), H13CO+ (1-0), and SiO (2-1), which is 1.51 × 106 cm−3, 1.83 × 105 cm−3, and 2.26 × 105 cm−3, respectively. We have shown that for a hydrogen density ≥1.0 × 105 cm−3, all these transitions are showing absorption features. If H2 density is below the critical density, level populations are mostly governed by collisional excitation and de-excitation rather than a Boltzmann distribution. Thus, these molecules might show emission even at T ≤ Tc. The observed spectral profile of SiO is in line with these results. We found SiO emission at all positions except the source's central region (where density is high, which is probably greater than the critical density).

Figure 5.

Figure 5. Variation of the radiation temperature of (a) HCN(1-0), (b) H13CO+ (1-0), and (c) SiO (2-1) respectively, for a wide range of parameter space. The vertical axis is the kinetic temperature, labeled Tkin (K). The units of the radiation temperature (Trad) used in the color bar are in Kelvin.

Standard image High-resolution image

3.5. Physical Properties of the Source

In this source, we obtained a temperature range 33–180 K. The hydrogen column density of this source is estimated as ∼ 1.53 × 1025 cm−2. Various observed molecular emissions are discussed below.

3.5.1. Infalling Envelope

Infall nature was observed previously in G31 by various authors (Girart et al. 2009; Mayen-Gijon et al. 2014; Beltrán et al. 2018, e.g.,). Girart et al. (2009) observed an inverse P-Cygni profile in G31 with low energy lines of C34S (7-6). Recently, Beltrán et al. (2018) identified redshifted absorption of H2CO and CH3CN and its isotopologues toward the dust continuum of the main core with different upper state energy. Mayen-Gijon et al. (2014) observed various inversion transitions (2,2), (3,3), (4,4), (5,5) and (6,6) of ammonia by using Very Large Array (VLA) observations to understand the infall motion in G31. To explain the infall motion in G31, they used a different approach called the central blue spot which can be easily identified in the first-order moment maps. The central blue spot consists of blueshifted emission in the first-order moment toward the zeroth order moment map's peak position. They found the central blue spot infall signature in all observed transitions of ammonia. In this work, we have detected one transition of H13CO+ (1-0) at frequency 86.75430 GHz, which shows both absorption and emission nature (see Figure 2(a)). This is an inverse P-Cygni profile. The infall velocity can be estimated as the difference between the systematic velocity and the velocity at the observed inverse P-Cygni profile's absorption feature. The infall velocity of the source is obtained to be 4.5 km s−1. This suggests that matter is infalling from the envelope toward the source's central region, which is consistent with the earlier observations (Girart et al. 2009; Beltrán et al. 2018).

3.5.2. Molecular Outflow

Many authors previously studied molecular outflows associated with G31 (e.g., Olmi et al. 1996; Cesaroni et al. 2011; Beltrán et al. 2018). SiO and 13CO revealed the presence of two molecular outflows: One in the east–west (E-W) direction and another one in the north–south (N-S) direction. SiO is an excellent tracer of shock and molecular outflow associated with star-forming regions (e.g., Zapata et al. 2009; Leurini et al. 2014). Beltrán et al. (2018) showed average redshifted and blueshifted wings of SiO emission which reveal three outflow directions associated with this source: E-W, N-S, and NE-SW. Among these three outflow directions, the N-S and E-W components are clearly present in our observation. However, the NE-SW component is not very evident (see Figure 6, left panel, of Beltrán et al. (2018) and Figure A1). Channel maps of SiO emission are shown in Figure 6. The blueshifted and redshifted integrated emission of SiO along with the continuum is depicted in Figure 7. The presence of HCN is widespread in diverse interstellar environments ranging from the dark cloud to the star-forming region and circumstellar envelope of a carbon-rich star (e.g., Snyder & Buhl 1971; Ziurys & Turner 1986). HCN also traces molecular outflow in G31. The observed spectrum profile of HCN (see Figure 2(b)) is symmetric with respect to the systematic velocity of the source. However, the intensity of emission wings associated with the absorption profile is asymmetric. This may be due to the asymmetry in molecular distributions. Figure 8 shows the redshifted and blueshifted integrated emissions of HCN overlaid on the continuum.

Figure 6.

Figure 6. Channel maps of SiO (2-1) emission. The ellipse at the lower left corner of the image shows the beam size of 1farcs52 × 1farcs09 with PA = 88fdg98. The white plus sign indicates the position of the continuum.

Standard image High-resolution image
Figure 7.

Figure 7. Gray scale levels for the continuum at 94 GHz. Blue-shifted emission of SiO integrated over 77–93 km s−1 where contour levels are at the 20, 30, 40, 50, 60, 70, 80% of the peak intensity. Redshifted emission of SiO is integrated over 100–123 km s−1. The ellipse at the lower left corner of the image shows the synthesized beam (1farcs19 × 0farcs98). The plus sign indicates the position of the continuum. The two crosses at the outer contour of the red lobes indicate lr1 and lr2 distance; the other two crosses on the red lobes indicate lb1 and lb2 distance.

Standard image High-resolution image
Figure 8.

Figure 8. Gray scale levels are used for the continuum at 94 GHz. Blue-shifted emission of HCN integrated over 77–90 km s−1 contours levels are at the 20, 30, 40, 50, 60, 70, and 80% of the peak intensity. Redshifted emission of HCN integrated over 105–115.7 km s−1. The ellipse at the lower left corner of the image shows the synthesized beam (1farcs19 × 0farcs98) with a PA = 76fdg83.

Standard image High-resolution image

To obtain the outflow parameters, we estimated the outflow velocity as the maximum velocity of outflow emission with respect to the systematic LSR velocity (97 km s−1). For the blue lobe and red lobe of SiO outflow, velocities are found to be 23 km s−1 and 33 km s−1. We estimated the dynamical timescale of outflow by using tdyn = $\tfrac{l(b,r)}{{V}_{\max }(b,r)}$. For this, we calculate two different lengths (lb1, lb2) of the blue lobe and two lengths (lr1, lr2) for red lobes (see Figures 78). We obtained an average dynamical timescale of the outflow of ∼ 4 × 103 years by considering the maximum velocity of SiO of the blue lobe and red lobe. The dynamical timescale of HCN outflow is found to be ∼3 × 103 yr. These values are in good agreement with the dynamical timescale of outflow of 12CO and SiO, as previously reported in G31 by Cesaroni et al. (2011) and Beltrán et al. (2018). For all individual positions, the different dynamical timescales for both blue and red lobes are presented in Table 5. For HCN, we calculate the outflow parameters for one blue lobe (lb1) and one red lobe (lr1). To estimate the other outflow parameters such as mass, momentum, energy, mass loss, momentum loss, and energy loss per year, we used the formula mentioned in Cabrit & Bertout (1992). To estimate the outflow parameters, we assumed the same abundance (1.0 × 10−8) and excitation temperature (20 K) of SiO (Codella et al. 2013) and HCN. All the outflow parameters are summarized in Table 5. All these outflow parameters are in good agreement with the observed values of Araya et al. (2008).

Table 5.  Physical Parameters of the Outflow

Parameters SiO HCN
Distance from continuum position to outer contour (pc)    
Red lobe (lr1) 0.117 (0.055) 0.118 (0.055)
Red lobe (lr2) 0.087 (0.041)
Blue lobe1 (lb1) 0.108 (0.051) 0.09 (0.042)
Blue lobe2 (lb2) 0.135 (0.063)
Typical outflow velocity: Voutflow (km s−1)    
Red lobe 33.0 40.0
Blue lobe 23.0 30.0
Dynamical time: tdyn (yr)    
Blue lobe 4.59 × 103 (1.62 × 103) 2.87 × 103 (1.35 × 103)
  5.74 × 103 (1.21 × 103)
Red lobe 3.47 × 103 (2.16 × 103) 2.90 × 103 (1.36 × 103)
  2.58 × 103 (2.70 × 103)
Average 4.10 × 103 (1.92 × 103) 2.88 × 103 (1.36 × 103)
Outflow mass (M)    
Blue lobe 27 (6) 12 (2.6)
Red lobe 24 (5) 55 (12)
Total 51 (11) 67 (14.6)
Momentum (M km s−1)    
Blue lobe 619 (136) 352 (77)
Red lobe 790 (173) 2213 (485)
Total 1409 (309) 2565 (562)
Kinetic energy (L yr)    
Blue lobe 7.1 × 103 (1.6 × 103) 5.3 × 103 (1.2 × 103)
Red lobe 1.3 × 104 (8.8 × 103) 4.4 × 104 (9.7 × 103)
Total 2.0 × 104 (1.0 × 104) 4.9 × 104 (1.1 × 104)
Mass loss (M yr−1)    
Blue lobe 5.2 × 10−3 (4.1 × 10−3) 4.1 × 10−3 (1.9 × 10−3)
Red lobe 7.9 × 10−3 (2.1 × 10−3) 1.9 × 10−2 (8.9 × 10−3)
Total 1.3 × 10−2 (6.2 × 10−3) 2.3 × 10−2 (10.8 × 10−3)
Momentum loss (M km s−1 yr−1)    
Blue lobe 0.12 (0.09) 0.12 (0.06)
Red lobe 0.26 (0.07) 0.76 (0.35)
Total 0.38 (0.16) 0.88 (0.41)
Energy loss (L)    
Blue lobe 1.38 (1.10) 1.84 (0.85)
Red lobe 4.31 (3.63) 15.26 (7.14)
Total 5.69 (4.73) 17.10 (7.99)

Note. To estimate the SiO outflow physical parameters, we considered an average value of the red lobe $\left({{r}}_{r}=\tfrac{{lr}1+{lr}2}{2}\right)$ and blue lobe $\left({{r}}_{{b}}=\tfrac{{lb}1+{lb}2}{2}\right)$ and the corresponding average timescales of the blue lobe and red lobe. We have calculated the lobe distance in arcseconds and then converted into parsecs by considering a source distance of 7.9 kpc. In parentheses, we have noted the outflow parameters by considering the new distance of 3.7 kpc for G31.

Download table as:  ASCIITypeset image

In Figure 6, we have shown the channel maps of SiO emission associated with the HMC. This figure depicts that blueshifted channels are distributed in the southwest region and redshifted channels toward the southeast and northeast regions. Bipolar lobe morphology observed from 93.2 km s−1 and 103.1 km s−1 channels suggests wide-angle bipolar outflow. Integrated emission of SiO and HCN transition (Figures 7 and 8) shows that there is a clear outflow direction along the east–west direction. In the integrated emission, the red lobe of the outflow is slightly tilted to the southeast. This may occur due to another outflow direction along the southwest-northeast direction. There is also an outflow in the N-S direction. Outflow of HCN along the N-S direction is wider (see Figure 8) as compared to the SiO outflow (see Figure 7). This may be attributed to the excitation conditions and chemical origin of HCN. The estimated outflow is massive, 50 M for SiO and 67 M for HCN and the outflow is characterized as young (3–4 × 103 yr). However, by considering the new distance measurement of ∼3.7 kpc of this source, we obtained a dynamical timescale of outflow around 1–2 × 103 yr. The estimated outflow mass is ∼11 M for SiO and ∼14.6 M for HCN.

3.6. Molecular Distributions and Abundances

In this section, we present the molecular abundance of all the observed species. We compare the abundance of all detected species between G31 and other sources, which is presented in Table 4. All the molecular abundances are reported with regard to H2. Figure A1 in the Appendix shows the schematic diagram of kinematics and molecular distributions of G31 based on the present observation. There are three outflow directions; one is along the E-W direction, one is along the N-S direction, and another is along the NE-SW direction (only for SiO). Outflow components are traced by HCN and SiO emissions. Here, we have obtained different molecular distributions of COMs. The emitting regions of different transitions of COMs are different. We have separated those into three regions such as the emitting region is smaller than the beam size (θ ∼ 0farcs8; see the purple circle in Figure A1), comparable to the beam size (θ ∼ 1farcs1; see the green circle outer edge in Figure A1), and greater than the beam size (θ ∼ 1farcs4; see the cyan circle in the Figure A1). The emitting regiona of all observed species are provided in Table B1 (see Appendix B). The uncertainties of emitting regions listed in Table B1 are the lower limit of true values as estimated emitting regions might be affected by several issues such as poor angular resolution, image noise, and the non-Gaussian shape of the source. Emitting regions of CH3OH and CH3OCHO are comparable or slightly greater as compared to the beam size. Molecular distributions show that CH3NCO and CH3SH have a slightly lower-emitting region as compared to beam size (marginally resolved). The data presented in this paper suggest possibly different emitting regions for different COMs, but these are not conclusive because of comparable beam and source size. High sensitivity data with high angular and spatial resolution is required to understand the spatial distributions of these molecules.

3.6.1. Sulfur Dioxide, SO2

SO2 is a good tracer of dense and hot gas of typical hot core (e.g., Charnley 1997; Leurini et al. 2007). In this source, we detected a strong transition (83,5 − 92,8) of SO2. Using LTE model, we estimated the column density and the fractional abundance of SO2, which are ∼2.88 × 1018 cm−2 and 1.88 × 10−7. The observed abundance of SO2 in G31 is one order of magnitude higher as compared to Srg B2, Orion KL and G10.47+0.03.

3.6.2. Formaldehyde, H2CO

The presence of H2CO with its ortho, para, and isotopic form within the frequency range of 218–364 GHz was previously reported by Isokoski et al. (2013). Redshifted strong absorption of H2CO (Eup = 21 K, 68 K) had recently been observed in G31, which supports the existence of infall in the core (Beltrán et al. 2018). Previously, Di Francesco et al. (2001) reported a similar kind of inverse P-Cygni profile and found that H2CO traces infall in IRAS 4A and they also detected the outflow from both the 4A and 4B positions by observing bright wing emission of H2CO. Formaldehyde, along with its deuterated counterpart, has also been observed in outflows toward other sources (e.g., Sahu et al. 2018). The presence of H2CS in G31 has previously been reported by Ligterink et al. (2015). Here, we also have detected one strong emission line of H2CS at 101.478 GHz. However, the observed spectrum of H2CS shows that there is a blending with other molecular lines (methyl formate, confirmed from LTE modeling). H2CO and H2CS are ubiquitous in the ISM. These two species have similar chemical structures except the O atom of H2CO is replaced by a S atom in H2CS.

3.6.3. Methanol, CH3OH

Methanol is a well known, and one of the most abundant, COMs in hot cores. Here, we have identified several transitions of methanol. The LTE model suggests that it has a high fractional abundance ∼1.2 × 10−6. We have also estimated methanol abundance following rotation diagram analysis. The obtained abundance using rotation diagram analysis is ∼1.2 × 10−6, which is similar to the estimated results using the LTE model. A comparison of the fractional abundance of methanol between G31 and other hot molecular cores is presented in Table 4.

3.6.4. Methanethiol, CH3SH

Methanethiol has a similar chemical structure as methanol except that the O atom of methanol is replaced by a S atom of methanethiol. In Figure A7, we show the observed and model (LTE) spectra of CH3SH. The best-fit abundance and column density of methanethiol are ∼2.70 × 10−8 and ∼4.13 × 10−17 respectively. We also estimated the column density and fractional abundance of this species using rotation diagram analysis, which is one order of magnitude lower compared to LTE model fitted results. The observed emitting region of methanethiol is comparable to the beam size. The spatial distribution of its emission is as compact as the dust continuum since our beam size is comparable to the source size. Therefore, we have uncertainties in determining the emitting regions of this species. For detailed information about the distribution of this species, we need high angular and spatial resolution data.

3.6.5. Methyl Formate, CH3OCHO

CH3OCHO is one of the most abundant COMs, which was observed in both low-mass and high-mass star-forming regions (e.g., Brown et al. 1975; Cazaux et al. 2003). Previously, numerous transitions of CH3OCHO were identified in G31 (Isokoski et al. 2013; Rivilla et al. 2017; Beltrán et al. 2018). Column density, fractional abundance, and a comparison of the fractional abundance of methyl formate with other hot cores are provided in Table 4. The estimated column density and fractional abundance of methyl formate are similar using both the rotation diagram and LTE model fitting analysis. Emitting diameters of various transitions of CH3OCHO are noted in Table B1.

3.6.6. Methyl Isocyanate, CH3NCO

In the case of CH3NCO, we have obtained a lower value of rotational temperature (48 K, see Figure 4(a)), which is low compared to a typical hot core. Methyl isocyanate transitions are compact and yield lower value rotational temperature. A similar trend has also been observed for the cold component of methanol. Figure A8 shows the observed and model (LTE) spectra of CH3NCO. The LTE model suggests that for a column density of 7.22 × 1017 cm−2, we have a good fit with the observed spectra. We have also estimated the column density of CH3NCO using rotation diagram analysis, which is one order higher compared to the LTE model fitting result. CH3NCO emissions are also concentrated toward the center of the source's dust continuum. However, all these transitions are marginally resolved. Thus, we cannot provide the detailed distributions of this molecule.

4. Discussion

Synthesis of many COMs (e.g., ethylene glycol, methyl formate) in G31 has been discussed by Rivilla et al. (2017). They discussed the formation routes of a few COMs. Thus, we do not concentrate on the synthesis of all the species which are identified in the present observation. Here, we mainly focus on the chemistry of two newly reported molecules, CH3NCO and CH3SH, and two other COMs, CH3OH and CH3OCHO, which are discussed in the text. Here, we have discussed the possible synthesis routes of these species and compared their observed abundance/molecular ratio between G31 and other hot molecular cores. We have also shown a comparison between observed results and the existing astrochemical simulation of these species. For the modeling results, we mainly followed Garrod (2013). Garrod considered an isothermal collapse phase followed by a static warm-up phase. In the first phase, the density increases from nH = 3 × 103 to 107 cm−3 under freefall collapse. At the initial phase, the dust temperature falls to 8 K from 16 K. In the second phase, the temperature varies from 8 K to 400 K, where density remains fixed at 107 cm−3. The temperature of this source is ∼150 K, which is typical hot core temperature, and the number density (nH) of this source is ∼107 cm−3 (Beltrán et al. 2018). Thus, the hot core model of Garrod (2013) is suitable for understanding the chemical evolution of COMs in this source. Garrod used three different warm-up models based on the timescale (i.e., fast, medium, and slow) and the final temperature of the model is much larger than that observed in this source. The timescale of the fast warm-up model is more suitable for studying the chemical evolution in high mass star-forming regions. We used peak gas phase abundances of all the species around 100–200 K of the fast warm-up timescale model to compare with the observed values. Figure 9 compares the observed fractional abundance of CH3OH, CH3SH, CH3OCHO, and CH3NCO between G31 and other sources. The chemistry of these COMs is discussed in the following subsections.

Figure 9.

Figure 9. Comparison of abundance of CH3OH, CH3SH, CH3CHO, and CH3NCO between G31 and other sources (Sgr B2, Orion KL, G10.47+0.03, and model). These abundances are given in Table 4 with their references.

Standard image High-resolution image

4.1. CH3OH and CH3SH

CH3OH and CH3SH molecules are the chemical analog of each other and have a similar structure; the only difference is that one has an oxygen atom and the other has a sulfur atom. The emitting regions of CH3SH transitions are compact and marginally resolved, whereas the emitting region of methanol is slightly extended compared to CH3SH. A large number of previous experimental, theoretical, and observed results suggest that these two species are mainly produced on the grain surface via successive hydrogen addition reactions (e.g., Das et al. 2008, 2010, 2016; Das & Chakrabarti 2011; Fedoseev et al. 2015; Müller et al. 2016; Gorai et al. 2017; Vidal et al. 2017) with CO and CS, respectively. The main route of formation of methanol and methanethiol is given below,

and

For the above reaction schemes, the first and third steps possess an activation barrier, whereas the second and fourth steps are barrier-less (e.g., Hasegawa et al. 1992; Gorai et al. 2017). At low-temperature regions, methanol and methanethiol are efficiently produced on the grain surface and released to the gas phase via non-thermal desorption in the cold phase and during the warm-up stage via a thermal desorption process. From the astrochemical simulation, Müller et al. (2016) found that the peak abundance of methanol occurs around 130 K. Here, our observed results of methanol predict a gas temperature of ∼180 K, which is similar to the typical hot core temperature. In addition to grain surface reactions, H2CS can also be produced in the gas phase. In the gas phase, H2CS is mainly produced from H3CS+, where H3CS+ is formed via the ion-neutral reaction between S+ and CH4. Below ∼20 K with a density (nH) around 104–105 cm−3, H2CS is accreted onto the grain surface directly from the gas phase, which may also contribute to the CH3SH formation (Bonfand et al. 2019).

The observed abundance of CH3SH in G31 is 2.7 × 10−08 and 1.0 × 10−09 in Orion KL (Kolesniková et al. 2014, see Table 4). The observed abundance of methanethiol in Sgr B2 is ∼3.5 times higher than the G31. Methanol abundance is almost similar in G31, Orion KL, G10.47+0.03, and 10 times higher in Sgr B2 (see Figure 9). The observed CH3SH/CH3OH ratio in G31 is similar to Orion KL and it is eight times higher in Sgr B2. Modeling results of these two species also have comparable abundances and molecular ratios (see Table 4).

4.2. CH3OCHO

Methyl formate can efficiently be produced on the surface of dust grains through the reaction between methoxy radical (CH3O) and formyl radical (HCO; Garrod et al. 2008; Garrod 2013). Garrod et al's simulation shows that, around 30–40 K, these radicals are mobile, and the reaction is efficient. It is clear from their simulation that gas phase CH3OCHO is mainly coming from the ice phase (see Figure 1 of Garrod 2013). Increased UV photodissociation of CH3OH leads to the formation of various radicals CH3, CH3O, and CH2O, around 40 K temperature; these are mobile and form the COMs (e.g., CH3OCHO, CH3OCH3). In the gas phase, when T ∼ 40 K, protonated methanol and H2CO (when it is abundantly released to the gas phase) react and form HC(OH)OCH${}_{3}^{+}$. Then electron recombination of HC(OH)OCH${}_{3}^{+}$ produces CH3OCHO (Bonfand et al. 2019). An efficient gas phase reaction of methyl formate production was proposed by Balucani et al. (2015) for the cold environment, where dimethyl ether is the precursor of methyl formate. Dimethyl ether is also present in this source (Isokoski et al. 2013). Therefore methyl formate can be produced through both grain surface and gas phase reactions. The observed abundance of CH3OCHO in G31 is 4.25 × 10−7, which is similar to the observed abundance in Sgr B2, G10.47+0.03, and 10 times higher as compared to the Orion KL and modeling results. In G31, we have observed a CH3OCHO/CH3OH abundance ratio of ∼0.14–0.35. For other high mass star-forming regions, this ratio is found to be 0.03, 0.03, and 0.08 in Sgr B2, Orion KL, and G10.47+0.03, respectively (see Table 4).

4.3. CH3NCO

In this work, we find the emitting regions of CH3NCO are compact. In this source, Rivilla et al. (2017) also observed a similar emitting region of ethylene glycol. It is primarily produced on the grain surface via recombination of two HCO radicals followed by successive hydrogen addition reactions (Coutens et al. 2017). In the case of CH3NCO formation, Halfen et al. (2015) proposed two routes: (i) a reaction between HNCO(HOCN) and CH3 and (ii) a reaction between HNCO(HOCN) and ${\mathrm{CH}}_{5}^{+}$. Cernicharo et al. (2016) proposed that it could be produced on the grain surface. The experiment result suggests that CH3NCO can efficiently be produced on the solid phase through the reaction between CH3 and (H)NCO (Ligterink et al. 2017). The hot core model suggests that it could be produced via a radical–radical reaction between CH3 and OCN (Belloche et al. 2017). Since the reaction mentioned above is exothermic and barrier-less, it can efficiently be produced on the grain surface at a little warmer temperature (>30 K), when radicals get enough mobility. Though CH3NCO is efficiently produced on the grain surface, it could also be destructed by hydrogen addition reactions and produce n-methyl formamide (Belloche et al. 2017). Quénard et al. (2018) also considered similar radical–radical reactions in their chemical model to study the evolution of CH3NCO under hot core conditions. In the hot core, when the temperature reaches around 100 K, CH3 and HNCO can thermally desorb from the grain surface and increase the formation of CH3NCO. Our estimated CH3NCO abundance in G31 is ∼4.72 × 10−8. For other hot molecular cores (e.g., Sgr B2, Orion KL), we have found a comparable abundance of CH3NCO. The observed fractional abundance of CH3NCO in this work is well within the modeled results (model 5) of Belloche et al. (2017).

The identification of CH3OH, CH3OCHO, CH3NCO, and CH3SH in G31 suggests both grain surface and gas phase chemistry are efficient for the formation of COMs in this source. However, we cannot distinguish the contribution from the gas and grain phase separately due to the observational limit. To explain the various observed line profiles in this work requires a combined chemical and radiative transfer model, which will be carried out in our follow-up study.

5. Conclusions

We analyzed ALMA band 3 data of a hot molecular core, G31.41+0.31, and the following conclusions are made.

  • 1.  
    We detected methyl isocyanate (CH3NCO) and methanethiol (CH3SH) for the first time in G31.
  • 2.  
    We obtained two temperature components by using rotation diagram analysis. The cold component is obtained by CH3OH and CH3NCO where temperature varies between 33 and 48 K. Hot component is obtained by CH3SH, CH3OH, and CH3OCHO where the temperature ranges between 143 and 181 K. The estimated hydrogen (${N}_{{{\rm{H}}}_{2}}$) column density is ∼1.53 × 1025 cm−2 and dust is optically thin at 94 GHz.
  • 3.  
    We estimated column density and fractional abundances of various complex species CH3NCO, CH3SH, CH3OH, and CH3OCHO in G31. A comparison of column density/fractional abundances of these species between G31 and other hot molecular cores is discussed. We also compared our observed results with available chemical modeling under hot core conditions. Our observed abundances of COMs in G31 are consistent with other hot cores, Sgr B2, Orion KL, and G10.47+0.03, which suggests similar formation mechanisms of these complex species in these sources.
  • 4.  
    For the kinematics of this source, a low excitation line of H13CO+ is observed as an inverse P-Cygni profile, which suggests that there is an infall toward the center of this source, i.e., this source is in the younger stage of its evolution.
  • 5.  
    Protostellar outflows are traced by SiO and HCN. Their blueshifted and redshifted emission clearly traces outflow along the east–west direction. We estimated various outflow parameters. The average dynamical timescale of outflow is ∼103 yr.

This paper makes use of the following ALMA data: ADS/JAO.ALMA#2015.1.01193.S. ALMA is a partnership of ESO (representing its member states), NSF (USA), and NINS (Japan), together with 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 ESO, AUI/NRAO, and NAOJ. We are grateful to the anonymous referee for numerous comments and suggestions that improved the manuscript. P.G. acknowledges a CSIR extended SRF fellowship (grant No. 09/904 (0013) 2018 EMR-I) and Chalmers Cosmic Origins postdoctoral fellowship. A.D. acknowledges the ISRO respond project (grant No. ISRO/RES/2/402/16-17) for partial financial support. B.B. acknowledges a DST-INSPIRE fellowship [IF170046] for providing partial financial assistance. S.K.M. acknowledges a CSIR fellowship (Ref no. 18/06/2017(i) EU-V). This work is partially supported by the Consortium for the Development of Human Resources in Science and Technology by the Japan Science and Technology Agency. This research was possible in part due to a Grant-in-Aid from the Higher Education Department of the Government of West Bengal. D.S. acknowledges an ASIAA postdoctoral research fund and travel grant. A.D. and T.S. also acknowledge support from a DST JSPS grant.

Software: CASA (v4.7.2; McMullin et al. 2007), CASSIS (Cassis Team At CESR/IRAP 2014), RADEX (Van der Tak et al. 2007).

Appendix A: Fitted Spectra, LTE Model Spectra, Moment Maps, and Schematic Diagram of G31 Based on the Present Observation

We show a schematic diagram of G31 based on the present observation in Figure A1, Gaussian fitted spectra in Figures A2A6, LTE model spectra in Figures A7 and A8, and moment maps in Figures A9A12.

Figure A1.

Figure A1. Schematic diagram (not up to scale) of kinematics and molecular distribution in G31.41+0.31 based on present observational results (see details in the Section 3.6.6).

Standard image High-resolution image
Figure A2.

Figure A2. The black line represents observed emission spectra of methanol (CH3OH) and the red line represents a Gaussian profile fitted to the observed spectra.

Standard image High-resolution image
Figure A3.

Figure A3. The black line represents observed emission spectra of methyl formate (CH3OCHO) and the red line represents a Gaussian profile fitted to the observed spectra. The 1139 − 112,10(A) transition may blend with some other molecular transition.

Standard image High-resolution image
Figure A4.

Figure A4. The black line represents observed emission spectra of methyl isocyanate (CH3NCO) and the red line represents a Gaussian profile fitted to the observed spectra. The 10−30–9−30 transition may strongly blend with some other molecular transitions.

Standard image High-resolution image
Figure A5.

Figure A5. The black line represents observed emission spectra of methanethiol (CH3SH) and the red line represents the resultant Gaussian profile of multiple Gaussian components fitted to the observed spectra. In the left panel, the blue profile represents the 101.13915 GHz line and the green line represents the 101.13965 GHz transition. In the middle panel, although we observe four transitions, we consider two Gaussian components, one (blue profile) for 101.15933 GHz and another one (green profile) for the other three multiplets (101.15999 GHz, 101.16066 GHz, and 101.16069 GHz). In the right panel, blue and green lines represent 101.16715 GHz and 101.16830 GHz transitions, respectively.

Standard image High-resolution image
Figure A6.

Figure A6. The black line represents observed emission spectra of SO2, and H2CO and the red line represents a Gaussian profile fitted to the observed spectra.

Standard image High-resolution image
Figure A7.

Figure A7. The black line represents observed spectra and the blue line represents the (LTE) model spectra of methanethiol (CH3SH). In addition to the observed transitions of CH3SH listed in Table 1, we have also seen signatures of two other transitions (101.17982 GHz and 101.20417 GHz), which are heavily blended with other species.

Standard image High-resolution image
Figure A8.

Figure A8. The black line represents observed spectra and the blue line represents the (LTE) model spectra of methyl isocyanate (CH3NCO).

Standard image High-resolution image
Figure A9.

Figure A9. Moment 0 maps of the integrated intensity distribution (color) of CH3OH transitions are overlaid on the 3.1 mm continuum emission (black contours). Contour levels are at 20%, 40%, 60%, and 80% of the peak flux of the continuum image. The synthesized beam is shown in the lower left-hand corner of each figure.

Standard image High-resolution image
Figure A10.

Figure A10. Moment 0 of the integrated intensity distribution (color) of CH3OCHO transitions is overlaid on the 3.1 mm continuum emission (black contours). Contour levels are at 20%, 40%, 60%, and 80% of the peak flux of the continuum image. The synthesized beam is shown in the lower left-hand corner of each figure.

Standard image High-resolution image
Figure A11.

Figure A11. Moment 0 of the integrated intensity distribution (color) of CH3NCO transitions is overlaid on the 3.1 mm continuum emission (black contours). Contour levels are at 20%, 40%, 60%, and 80% of the peak flux of the continuum image. The synthesized beam is shown in the lower left-hand corner of each figure.

Standard image High-resolution image
Figure A12.

Figure A12. Moment 0 of the integrated intensity distribution (color) of CH3SH transitions is overlaid on the 3.1 mm continuum emission (black contours). Contour levels are at 20%, 40%, 60%, and 80% of the peak flux of the continuum image. The synthesized beam is shown in the lower left-hand corner of each figure. Multiple transitions (101.13915/101.13965 GHz) appear in CH3SH 40,4-30,3 moment maps. Similarly multiple transitions (101.16715/101.16830 GHz) appear in CH3SH 42,2-32,1 moment maps.

Standard image High-resolution image

Appendix B: Summary of the Emitting Regions of the Observed Species

We list in Table B1 the summary of the emitting regions of the observed species.

Table B1.  Emitting Region of Observed Transitions of COMs

Species Transition Emitting Upper State
  Frequency Region Energy
  (GHz) ('') (K)
CH3SH 101.13915a 0.89 ± 0.06 12.14
  101.15999b 0.80 ± 0.02 52.39
  101.16830c 0.80 ± 0.05 30.27
       
CH3OH 86.61557 1.35 ± 0.01 102.70
  86.90291 1.35 ± 0.02 102.72
  88.59478 1.29 ± 0.02 328.28
  88.93997 1.27 ± 0.01 328.28
  101.12685 1.14 ± 0.03 60.73
  101.29341 1.21 ± 0.02 90.91
  101.46980 1.23 ± 0.03 109.49
       
CH3NCO 86.68019 0.82 ± 0.01 22.82
  86.68655 0.83 ± 0.02 34.97
  86.78077 0.78 ± 0.02 46.73
  86.80502 0.77 ± 0.01 46.74
  87.01607 0.76 ± 0.02 58.88
       
CH3OCHO 88.6869 1.37 ± 0.04 44.97
  88.72326 1.10 ± 0.03 44.96
  88.84318 1.36 ± 0.02 17.96
  88.85160 1.42 ± 0.02 17.94
  88.86241 1.15 ± 0.04 207.10
  101.37054 1.46 ± 0.05 59.64
  101.41474 1.33 ± 0.06 59.63
       
SO2 86.63908 0.71 ± 0.01 55.20
H2CO 101.33299 1.31 ± 0.02 87.57

Notes.

aMultiple transitions: 101.13915/101.13965 GHz. bMultiple transitions: 101.15999/101.15933/101.16066/101.16069 GHz. cMultiple transitions: 101.16715/101.16830 GHz.

Download table as:  ASCIITypeset image

Footnotes

Please wait… references are loading.
10.3847/1538-4357/abc9c4