Paper

Evolutions of CH3CN abundance in molecular clumps

, and

© 2021 National Astronomical Observatories, CAS and IOP Publishing Ltd.
, , Citation Zhen-Zhen He et al 2021 Res. Astron. Astrophys. 21 207 DOI 10.1088/1674-4527/21/8/207

1674-4527/21/8/207

Abstract

To investigate the effects of massive star evolution on surrounding molecules, we select nine massive clumps previously observed with the Atacama Pathfinder Experiment (APEX) telescope and the Submillimeter Array (SMA) telescope. Based on the observations of APEX, we obtain luminosity to mass ratios Lclump/Mclump that range from 10 to 154 L/M, where some of them embedded ultra compact (UC) HII region. Using the SMA, CH3CN (12K–11K) transitions were observed toward nine massive star-forming regions. We derive the CH3CN rotational temperature and column density using the XCLASS program, and calculate its fractional abundance. We find that CH3CN temperature seems to increase with the increase of Lclump/Mclump when the ratio is between 10 to 40 L/M, then decrease when Lclump/Mclump ≥ 40 L/M. Assuming that the CH3CN gas is heated by radiation from the central star, the effective distance of CH3CN relative to the central star is estimated. The distance ranges from ∼ 0.003 to ∼ 0.083 pc, which accounts for ∼ 1/100 to ∼ 1/1000 of clump size. The effective distance increases slightly as Lclump/Mclump increases (Reff ∼ (Lclump/Mclump)0.5±0.2). Overall, the CH3CN abundance is found to decrease as the clumps evolve, e.g., XCH3CN ∼ (Lclump/Mclump)−1.0 ± 0.7. The steady decline of CH3CN abundance as the clumps evolution can be interpreted as a result of photodissociation.

Export citation and abstract BibTeX RIS

1. Introduction

Massive stars (M ≥ 8 M) are formed inside molecular clouds. Feedback from star formation has substantial impacts on the evolution of the surround interstellar medium (ISM) through outflows, winds, as well as strong UV radiation (Garay & Lizano 1999; Zinnecker & Yorke 2007). Molecules, which have been detected in the ISM or circumstellar shells, are powerful tools to probe physical conditions such as densities, gas temperatures, and kinematical properties (Herbst & van Dishoeck 2009).

Gerner et al. (2014) observed a sample of high-mass star-forming regions at different evolutionary stages with the IRAM 30 m telescope, those stages varying from infrared dark clouds (IRDCs) to high-mass protostellar objects (HMPOs) to hot molecular cores (HMCs) and, finally, ultra compact (UC) H II regions. They found that molecular species are rich in the HMC phase and decline for the UC H II stage. However, the evolution of high-mass star formation occurs on a short time scale and in a clustered environment (Bonnell et al. 2003). The transition from one stage to the next is not well-defined, and there are some overlaps among those stages.

The collapse of molecular clump 1 is a global process: although the star formation that we observe occurs mostly at the center of the clumps, the actual collapse occurs on a much larger scale (e.g., Murray & Chang 2015; Li 2018). To study this collapse and the associated star formation, it is necessary to correlate the tracers of star formation with the properties of their natal clumps. A promising evolutionary indicator for massive and dense clumps is the L/M ratio – the ratio between the bolometric luminosity and the mass of the clump. Single dish investigations about the physical properties of star-forming regions at different L/M have been conducted in some works using molecules as probes (e.g., CH3C2H, CH3CN, NH3, CH3OH; Molinari et al. 2016; Giannetti et al. 2017). These works show that the molecular temperatures are significantly changed in different physical conditions during stars formation.

CH3CN (methyl cyanide) is known as a good tracer of warm and dense gas (e.g., Pankonin et al. 2001; Araya et al. 2005a; Giannetti et al. 2017). High spatial resolution observations seem to indicate that this molecule traces the disks around forming stars (Cesaroni et al. 1997; Keto & Zhang 2010; Cesaroni et al. 2014; Chen et al. 2016; Sanna et al. 2019). Interferometric observationswith high spatial resolutions are suitable for studying small regions as the CH3CN emission mostly originates from small scales. Hernández-Hernández et al. (2014) searched the literature for massive star-forming regions (MSFRs) in the HMC phase observed using Submillimeter Array (SMA) and selected 17 sources. They studied the chemical properties of CH3CN, and suggested that CH3CN was formed in gas phase and their abundance was increased with the increase of temperature. In this paper, we tend to use the sources selected by Hernández-Hernández et al. (2014) and place these sources in the context of star evolution to study the influence of star formation on the surrounding molecules. Following Molinari et al. (2008), we propose to study star evolutional stages indicated by the luminosity to mass ratios of the clumps.

2. Data

2.1. Sample

Our samples are taken from Hernández-Hernández et al. (2014), who searched the literature for MSFRs in the HMC phase. Most of their sample sources are associated with UC HII regions. These sources were observed at a spatial scale of ∼0.1 pc with similar observational resolution. We matched them with Atacama Pathfinder Experiment (APEX) Telescope Large Area Survey of the Galaxy (ATLASGAL; Schuller et al. 2009) dense clump catalogue based on their positions. The survey aims to provide a large and systematic list of clumps in the early embedded evolution stage. Nine massive clumps G5.89–0.39, G8.68–0.37, G10.62–0.38, G28.20–0.04N, G45.07+0.13, G45.47+0.05, IRAS 16547–4247, IRAS 18182–1433 and IRAS 18566+0408 are selected after removing some non-matched sources. The clump distances, effective radii, masses, and bolometric luminosities are taken from Urquhart et al. (2018). For our purpose, we use Lclump/Mclump – the ratio between clump luminosity and mass – as the evolutionary indicator, and the nine targets have Lclump/Mclump ranging from 10 to 154 L/M. A summary of clump properties is presented in Table 1.

Table 1. Clump Properties

SourceDistanceMassLuminosity Lclump/Mclump Rclump
 (kpc)(M)(105 L)(L/M)(pc)
G5.89–0.392.9916982.111240.653
G8.68–0.374.4522180.23100.712
G10.62–0.384.9586495.21601.584
G28.20–0.04N6.0544461.30291.349
G45.07+0.138.0031114.821541.357
G45.47+0.058.4071614.30592.078
IRAS 16547–42472.7416780.60350.691
IRAS 18182–14334.7113420.18130.708
IRAS 18566+04084.8619400.23111.084

Notes: Properties of our sample clumps were taken by matching them to the sample of ATLASGAL (Urquhart et al. 2018).

2.2. SMA Observations and Data Reduction

The SMA observations were performed between April 2004 and January 2009. Selected targets are listed in Table 2, where observational epoch, frequency range, and spectral resolution are included. The correlator was set to the double-sideband receiver — the lower sideband (LSB) and the upper sideband (USB), each bandwidth is about 2 GHz. The rest frequencies of CH3CN (12K–11K) are covered in the LSB. Spectral resolution is 0.406 MHz (∼0.53 km s−1) or 0.812 MHz (∼1.1 km s−1) for different sources. The system temperatures during those observations were less than 600 K.

Table 2. Observational Parameters

SourceObservationFrequency RangeSpectralCalibratorsSynthesized beam
 EpochLSBUSBResolutionBandpassGainFluxFWHMP.A.
  (GHz)(GHz)(MHz)   (arcsec)(deg)
G5.89–0.392008 Apr 18219.37–221.34229.37–231.340.4063C273nrao530 1921–293uranus2.65 2.18+54.0
G8.68–0.372008 Sep 17220.28–222.27230.28–232.270.4063C454nrao530 1911–201uranus7.86 1.99+16.4
G10.62–0.382009 Jan 31220.32–222.30230.32–232.300.4063C4541733–130 ...uranus5.37 2.66+9.2
G28.20–0.04N2008 Jun 21220.25–222.22230.25–232.220.4063C4541733–130 1911–201uranus2.77 2.59+50.8
G45.07+0.132007 Apr 13219.46–221.43229.46–231.430.8123C2791751+096 1925+211callisto2.87 1.65+77.5
G45.47+0.052008 Jun 30219.15–221.13229.15–231.130.4063C2791830+063 1925+211callisto3.11 2.69+71.4
IRAS 16547–42472006 Jun 06219.21–221.19229.21–231.190.8123C2731745–290 1604–446uranus1.83 0.96–7.5
IRAS 18182–14332004 Apr 30219.37–221.34229.37–231.340.8123C273nrao530 1908–201uranus3.19 2.00+16.5
IRAS 18566+04082007 Jul 09219.31–221.29229.31–231.290.8123C2731751+096 1830+063uranus3.06 1.48+60.6

Calibration and imaging are performed using the MIRIAD package 2 (Sault et al. 1995). Calibrators of bandpass, gain, and flux for each target are listed in Table 2. Considering the SMA monitoring of quasars, we estimate the flux uncertainty of ∼20%. We combined the continuum data from LSB and USB, the synthesized beam sizes range from 1.83'' × 0.96'' to 5.37'' × 2.66''. All the line data was regrided to a uniform spectral resolution of 0.812 MHz (∼1.1 km s−1).

3. Results and Analysis

Hernández-Hernández et al. (2014) had completed the analysis of those data, they derived CH3CN temperature and column density using a two-component method — a hot dense component and a warm extended component — by assuming different source sizes. In their results, the molecular abundances are mainly contributed from the hot dense components. In order to study the global variation of molecular abundance in star formation, we reanalyzed selected data using single component fitting, the source size was derived by fitting CH3CN (122–112) line emission region of each target (see Sect. 3.2), this method has been used in previous analysis (Beuther et al. 2006).

3.1. Continuum

The continuum images, as shown in Figure 1, are constructed from line-free channels of LSB and USB in the visibility domain. The 1σ noise level of continuum is lower than 20 mJy beam−1. Two-dimensional Gaussian fittings are performed to obtain peak intensities, total flux densities, and deconvolved source sizes. The relevant results are listed in Table 3.

Fig. 1

Fig. 1 1.3 mm continuum images (color scale) and molecules line emission (white contours) of CH3CN (122–112) toward massive star-forming regions. Contour levels have steps of 30%, 40%, 50%, 60%, 70%, 80%, until 90% of integrated emission. The synthesized beam of each source is shown at the bottom left. The 1.3 mm integrated emission flux of G5.89–0.39, G8.68–0.37, G10.62–0.38, G28.20–0.04N, G45.07+0.13, G45.47+0.05, IRAS 16547–4247, IRAS 18182–1433 and IRAS 18566+0408 is 7.50 Jy, 0.38 Jy, 6.86 Jy, 1.15 Jy, 1.63 Jy, 0.60 Jy, 0.56 Jy, 0.47 Jy, and 0.31 Jy, respectively, which are listed in Table 3.

Standard image

Table 3. Derived Parameters

Source1.3 mm Continuum ResultsDerived CH3CN parameters
${S}_{\nu }^{{\rm{Peak}}}$ ${S}_{\nu }^{{\rm{Total}}}$ ${S}_{\nu }^{{\rm{Dust}}\,1}$ ${\theta }_{{\rm{s}}}^{2}$ rms ${N}_{{\rm{H}}2}^{3}$ ${\theta }_{{\rm{m}}}^{4}$ rms Trot Ntot X ${R}_{{\rm{eff}}}^{5}$
(Jy beam−1)(Jy)(Jy)(arcsec)(mJy beam−1)(1024 cm−2)(arcsec)(mJy beam−1)(K)(cm−2) (pc)
G5.89–0.392.23±0.077.50±0.290.91±0.463.90×3.2411.260.705.439.02 ${147}_{-12}^{+17}$ ${8.63}_{-1.0}^{+1.3}$(14)1.23±0.78(-9)0.016±0.007
G8.68–0.370.22±0.010.38±0.010.25±0.013.67×2.200.580.391.215.65 ${115}_{-17}^{+27}$ ${1.88}_{-0.6}^{+0.5}$(15)4.82±3.69(-9)0.009±0.003
G10.62–0.382.98±0.126.86±0.373.78±1.024.33×2.963.212.162.415.47 ${194}_{-24}^{+49}$ ${2.51}_{-0.6}^{+0.8}$(15)1.16±0.90(-9)0.014±0.006
G28.20–0.04N0.87±0.011.15±0.010.62±0.211.73×1.121.501.881.431.50 ${240}_{-53}^{+46}$ ${6.23}_{-3.6}^{+2.5}$(15)3.31±3.25(-9)0.005±0.002
G45.07+0.131.13±0.031.63±0.070.73±0.421.31×1.184.094.731.331.14 ${143}_{-38}^{+15}$ ${2.89}_{-1.4}^{+0.3}$(15)6.10±4.74(-10)0.026±0.011
G45.47+0.050.42±0.010.60±0.020.35±0.092.02×1.552.182.151.530.12 ${77}_{-6}^{+13}$ ${4.34}_{-0.9}^{+1.2}$(14)2.01±1.47(-10)0.083±0.037
IRAS 16547–42470.24±0.010.56±0.030.43±0.071.39×1.2519.811.471.993.34 ${237}_{-30}^{+31}$ ${2.24}_{-0.7}^{+1.3}$(15)1.52±1.44(-9)0.003±0.001
IRAS 18182–14330.28±0.010.47±0.010.46±0.012.37×1.171.091.641.414.22 ${145}_{-22}^{+16}$ ${1.12}_{-0.2}^{+0.3}$(15)6.82±4.63(-10)0.005±0.002
IRAS 18566+04080.16±0.010.31±0.020.30±0.012.10×1.802.080.581.228.85 ${193}_{-25}^{+37}$ ${7.25}_{-2.9}^{+1.4}$(15)1.25±0.97(-8)0.003±0.001

Notes: 1 ${S}_{\nu }^{{\rm{Dust}}}={S}_{\nu }^{{\rm{Total}}}$${S}_{\nu }^{{\rm{free}}-{\rm{free}}}$. The free-free emission was obtained from Hernández-Hernández et al. (2014), who assumed the free-free emission is optically thin from 10 to 100 GHz and estimated the free-free emission by Sν ν−0.1. 2 Deconvolved continuum sizes from 2D Gaussian fit. 3 H2 column density derived from the estimated 1.3 mm dust continuum emission (${S}_{\nu }^{{\rm{dust}}}$) assuming Tdust = TCH3CN. 4 Deconvolved molecular distribution sizes from 2D Gaussian fit. 5 Effective distance of the CH3CN-emitting region is calculated using L = 4πR2 σT4. T is the temperature of the CH3CN gas and L is clump bolometric luminosity.

The 1.3 mm continuum emission may be contributed by free-free radiation and dust thermal radiation, because most HMCs harbor embedded UC H II regions. The free-free emission was obtained from Hernández-Hernández et al. (2014), who assumed the free-free emission is optically thin from 10 to 100 GHz and estimated the free-free emission by Sν ν−0.1. The estimated dust continuum flux of each source is listed in Table 3.

To derive CH3CN fractional abundance, we first calculate H2 column densities for all observed regions. Assuming that the dust emission is optically thin, H2 column density can be obtained by (Hildebrand 1983; Lis et al. 1991):

Equation (1)

where k and h are the Boltzmann constant and Planck constant, respectively, T is the dust temperature, Q(ν) is the grain emissivity at frequency ν, Ω is the solid angle of source, Sν is the total integrated flux of the dust continuum. We adopt Q(ν) = 2.2 × 10−5 at 1.3 mm (β = 1.5; Lis & Goldsmith 1990; Lis et al. 1991), and the gas-to-dust ratio of 100 is used. The dust temperature can be estimated from CH3CN rotational temperature, which will be derived in Section 3.2, assuming that dust and gas are in thermal equilibrium (Kaufman et al. 1998). The derived H2 column densities range from 0.39 × 1024 to 4.73 × 1024 cm−2, the values are listed in Table 3.

Considering an absolute flux uncertainty of ∼20% for SMA observations, dust temperature uncertainty of ∼20%, we estimate H2 column density uncertainties of ∼50% (Sanhueza et al. 2019). The H2 column density that we derived is comparable to the result of Hernández-Hernández et al. (2014). The difference may be caused by the fact that they used the CH3CN temperature of hot dense component to assume the dust temperature, while we derived a relatively low CH3CN temperature using single component fitting. The H2 column density is sensitive to dust temperature.

3.2. Molecular Line Emission

The spectral lines are extracted from the continuum peak position. Their continuum-subtracted spectra are plotted with intensity in units of Kelvin in Figure 2. The CH3CN (12K–11K) lines are identified following Hernández-Hernández et al. (2014).

Fig. 2

Fig. 2 Observed spectra (black) and model spectra (red) of CH3CN (12K–11K). The model spectra are derived with the XCLASS package. The number represent the K-ladder quantum numbers. The K = 9 line is overlapped with 13CO (2–1) in some regions.

Standard image

Spectral lines are modelled using the XCLASS package 3 (Möller et al. 2017) to derive molecular temperature and column density. The modeling parameters are source size, rotational temperature, column density, line width, and velocity offset. The source sizes are obtained by two-dimension Gaussian fits to CH3CN (122–112) line images. The 1σ noise level in line images is lower than 100 mJy beam−1. The line width and velocity offsets are obtained by Gaussian fitting. Those parameters are used as initial parameters. The optimization algorithm in MAGIX is used to explore the parameter space and minimize the χ2 distribution space. The detailed fitting functions and modelling procedures are described in Möller et al. (2017), where local thermodynamical equilibrium (LTE) is assumed. The derived molecular temperature and column density are presented in Table 3.

We derive molecular fractional abundances relative to H2 using X = Ntot/NH2 , where NH2 is the H2 column density. CH3CN rotational temperatures, column densities, and abundances are listed in Table 3. Both temperature and column density of CH3CN obtained by our single component fitting are lower than yet comparable to that of the hot dense component and higher than that of the warm extended component obtained by Hernández-Hernández et al. (2014).

4. Discussion

4.1. Evolutionary Stages

The distribution of ATLASGAL sources in the luminosity-mass (Lclump-Mclump) plane are presented in Figure 3. This type of diagram has been used in the studies of low-mass and high-mass star-forming regions (e.g., Saraceno et al. 1996; Molinari et al. 2008; Giannetti et al. 2017), and can be used as a tool for separating different evolutionary stages. Evolutionary tracks were derived by Molinari et al. (2008), vertical and horizontal arrow refer to as the accretion and the envelope dispersion phases. Three lines of constant Lclump/Mclump ratios (i.e., 1, 10 and 100 L/M) are shown in figure. Our sources at different evolutionary stages are marked in Figure 3.

Fig. 3

Fig. 3 Distribution of sources in the luminosity-mass plane. Stars represent selected sources in this paper. Gray dots represent the ATLASGAL clumps from Urquhart et al. (2018). The lower, middle and upper diagonal dash-dotted lines indicate the Lclump/Mclump = 1, 10 and 100 L/M, respectively. Vertical and horizontal arrows refer to as the accretion and the envelope dispersion phases derived by Molinari et al. (2008).

Standard image

Based on previous studies, G8.68–0.37, IRAS 18182–1433, and IRAS 18566+0408 have no or weak continuum emission at centimeter observations (Longmore et al. 2011; Beuther et al. 2006; Araya et al. 2005b), and they are massive star-forming regions at an early evolutionary stage before forming a significant UC H II region. Clumps reaching Lclump/Mclump ∼ 10 L/M are likely to stay in the transition phase between the main accretion phase and the dispersion phase (Giannetti et al. 2017). Most sources with Lclump/Mclump ≳ 30 L/M are harboring UC H II regions (Su et al. 2009; Hunter et al. 1997; Afflerbach et al. 1994; Wood & Churchwell 1989a), especially G5.89–0.39 and G45.07+0.13 have a Lclump/Mclump in excess of 100 L/M. G5.89–0.39 is an expanding shell-like UC H II region (Wood & Churchwell 1989b). Recently, Zapata et al. (2019) have presented sensitive CO(J = 32) observations that revealed the possible presence of an explosive outflow in G5.89–0.39. G45.07+0.13 is a pair of spherical UC H II regions, these regions are interpreted as a shell generated by stellar wind (Turner & Matthews 1984). A detailed description of each source is given in Hernández-Hernández et al. (2014).

Our sources have similar masses yet different luminosity-to-mass ratios. Taking advantage of this, we propose that our sources G8.68–0.37, IRAS 18566+0408, IRAS 18182–1433, G28.200.04N, IRAS 16547–4247, G45.47+0.05, G10.62–0.38, G5.89–0.39, and G45.07+0.13 form a sequence with Lclump/Mclump ranging from 10 to 154 L/M.

4.2. Temperature

In the upper left panel of Figure 4, we show the rotational temperature traced by CH3CN gas as a function of the Lclump/Mclump. It seems that the temperature increases with Lclump/Mclump when the ratio between 10 to 40 L/M, and decreases when Lclump/Mclump is greater than 40 L/M. It is puzzling that rotational temperatures do not demonstrate a monotonic increase over the increase of L/M ratios. One would expect, in an idealized situation, that as a massive protostar gains masses and luminosity, the protostellar heating raises temperatures of the surrounding gas. Therefore, rotational temperatures are expected to increase with L/M ratios. This trend was observed by single dish observations (e.g., Molinari et al. 2016; Giannetti et al. 2017; Urquhart et al. 2018). In these literatures, all molecular temperatures are considered to be continuously heated up when Lclump/Mclump ≳ 10 L/M.

Fig. 4

Fig. 4  Upper left: CH3CN rotational temperature as a function of the Lclump/Mclump. Molecular temperatures are estimated with the XCLASS package. Upper right: CH3CN abundance as a function of the Lclump/Mclump, the abundances are derived relative to H2 column density. Lower left: Derived size ratios between CH3CN effective distance Reff as a function of the Lclump/Mclump. Molecular effective distance Reff is derived based on molecular temperature and clump bolometric luminosity using L = 4πR2 σT4. Lower right: CH3CN abundance as a function of the ionization flux. The ionization flux at 1.3 mm was derived by Hernández-Hernández et al. (2014), who search for reported 10 to 100 GHz emission from the literatures and following Sν ν−0.1.

Standard image

Several factors may bias the data. The most direct influence on the observation results is the spatial resolution. The spatial resolution range of our nine sources is from ∼10000 to ∼20000 AU, except for IRAS 16547–4247 with ∼3000 AU. The Lclump/Mclump of IRAS 16547–4247 is 35 L/M, and it cannot result the decreasing trend of molecular temperature when Lclump/Mclump > 40 L/M. Thus the spatial dilution effect is not significant. Therefore, when Lclump/Mclump > 40 L/M, the decreasing trend of temperature is not caused by spatial resolution. The expansion of embedded UC H II regions might also affect the gas temperature. The Lclump/Mclump ∼ 10 L/M is considered a threshold that the birth of the zero age main sequence (Urquhart et al. 2014; Molinari et al. 2016; Giannetti et al. 2017), where hydrogen burning begins, then UC H II regions start to form and disperse the parent clumps (Hoare et al. 2007). Those processes may push ISM into the outer warm envelope, resulting a cavity located in the molecular clouds (e.g., G5.89, see Fig. 1). Such an expansion of UC H II region may cause molecular temperature decrease when Lclump/Mclump ≥ 40 L/M, where dissipation may dominate in this phase (Giannetti et al. 2017).

4.3. Location of CH3CN-emitting Region

In Figure 1, the CH3CN emission coincides with the 1.3 mm continuum emission, viewed at our current resolution. At the same time, we noticed that CH3CN has a relatively complicated morphology in G5.89, where multiple dust continuum regions were found (Hunter et al. 2008).

Assuming that the gas is heated by the central star, molecular temperature is determined by radiative heating, molecular effective distance relative to the central heating source can be estimated by:

Equation (2)

where R is the distance of the region with temperature T, the central source has bolometric luminosity L. We use this formula to estimate molecular effective distance Reff, based on the bolometric luminosity of source, and on the molecular temperature derived in Section 3.2. Keto et al. (1987) studied the temperature scales of NH3 gas in the vicinity of UC H II region G10.6–0.4, where the NH3 gas temperature scales outward with a radius as R−1/2.

The molecular effective distances were estimated using Equation (2) ranging from ∼ 0.003 to ∼ 0.083 pc, shown in Table 3. It is likely that CH3CN originated from circum-stellar instead of interstellar. We also plot the relation between Reff and Lclump/Mclump in the lower left panel of Figure 4. According to our estimation, CH3CN radiation should originate from ∼ 1/100 to ∼ 1/1000 of the ATLASGAL clump size and there is evidence for a correlation between the Reff and Lclump/Mclump where Reff ∼ (Lclump/Mclump)0.5 ± 0.2. As the star evolves, the molecule is located relatively farther away from the central star.

4.4. CH3CN Abundance as an Evolutionary Tracer

In the upper right panel of Figure 4, we plot the relation between the molecular abundance and the Lclump/Mclump. It seems that the abundance decreases as Lclump/Mclump increases, where a correlation of XCH3CN ∼ (Lclump/Mclump)−1.0 ± 0.7 can be found. CH3CN is abundant in the early stage of HMC, and then the abundance decreases with the formation and evolution of UC H II region.

As stars evolve, they release more and more energy through UV radiation. Chemical model revealed that molecules have high column density at lower UV fields, while many species are photodissociated away for high UV fields (Stäuber et al. 2004). CH3CN is subject to such photodissociation by UV radiation. The species whose abundances are enhanced are simple molecules including radicals and ions in photodissociated region (PDR), there are strong lines of simple species containing 2–4 atoms in W3 IRS4 (Helmich & van Dishoeck 1997), supporting the photodissociation hypothesis. As shown in the bottom right of Figure 4, molecular abundance decreases with the increase of ionization flux (XCH3CN ∼ Ionization flux−0.3±0.2). It can also support the explanation of the decrease of molecular abundance caused by ionization.

5. Conclusion

We present SMA observations of nine massive star-forming regions with Lclump/Mclump ratios ranging from 10 to 154 L/M. We detect CH3CN (12K–11K) lines in all sources, derive molecular temperatures and abundances, and study the relation between CH3CN abundances and the evolutionary stage of sources measured in terms of luminosity-to-mass ratio Lclump/Mclump.

We find that the rotational temperatures of CH3CN increase with the increase of Lclump/Mclump when the ratio is between 10 to 40 L/M, then seem to decrease when Lclump/Mclump ≥ 40 L/M, where the decline can be explained by dissipation. Assuming that the CH3CN molecules are heated by radiation from the central stars, we estimated the effective distance of the CH3CN relative to the central heating sources. Estimated CH3CN effective distance range from ∼ 0.003 to ∼ 0.083 pc, which accounts for ∼ 1/100 to ∼ 1/1000 of clump size. The molecular effective distance Reff increases slightly as Lclump/Mclump increases (Reff ∼ (Lclump/Mclump)0.5 ± 0.2).

We also find that the CH3CN abundance is anti-correlated with Lclump/Mclump where XCH3CN ∼ (Lclump/Mclump)−1.0 ± 0.7. They can be explained by photodissociation. CH3CN abundance decrease when ionization flux was high. The strong anti-correlation deserves to be investigated with future high angular-resolution observations, and the relative abundance of CH3CN might serve as a tracer for evolution in future studies.

Acknowledgements

Guang-Xing Li acknowledges support from the National Natural Science Foundation of China (Grant Nos. W820301904 and 12033005).

Footnotes

Please wait… references are loading.
10.1088/1674-4527/21/8/207