This site uses cookies. By continuing to use this site you agree to our use of cookies. To find out more, see our Privacy and Cookies policy.

THE DISK-OUTFLOW SYSTEM IN THE S255IR AREA OF HIGH-MASS STAR FORMATION

, , , , , , , , , and

Published 2015 August 24 © 2015. The American Astronomical Society. All rights reserved.
, , Citation I. Zinchenko et al 2015 ApJ 810 10 DOI 10.1088/0004-637X/810/1/10

0004-637X/810/1/10

ABSTRACT

We report the results of our observations of the S255IR area with the Submillimeter Array (SMA) at 1.3 mm in the very extended configuration and at 0.8 mm in the compact configuration as well as with the IRAM 30 m at 0.8 mm. The best achieved angular resolution is about 0.4 arcsec. The dust continuum emission and several tens of molecular spectral lines are observed. The majority of the lines is detected only toward the S255IR-SMA1 clump, which represents a rotating structure (probably a disk) around the young massive star. The achieved angular resolution is still insufficient to make any conclusions about the Keplerian or non-Keplerian character of the rotation. The temperature of the molecular gas reaches 130–180 K. The size of the clump is about 500 AU. The clump is strongly fragmented as follows from the low beam-filling factor. The mass of the hot gas is significantly lower than the mass of the central star. A strong DCN emission near the center of the hot core most probably indicates a presence of a relatively cold (≲80 K) and rather massive clump there. High-velocity emission is observed in the CO line as well as in lines of high-density tracers HCN, HCO+, CS and other molecules. The outflow morphology obtained from a combination of the SMA and IRAM 30 m data is significantly different from that derived from the SMA data alone. The CO emission detected with the SMA traces only one boundary of the outflow. The outflow is most probably driven by jet bow shocks created by episodic ejections from the center. We detected a dense high velocity clump associated apparently with one of the bow shocks. The outflow strongly affects the chemical composition of the surrounding medium.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

The formation of high mass stars (more massive than 8–10 M) is still poorly understood. Studies of this process are currently a "hot topic" of astrophysical research. A major unsolved problem of high-mass star formation is the characterization of accretion disks around young high-mass protostars. To date, most of evidence for the existence of such disks has been indirect (e.g., Beuther et al. 2009). If present, these disks undoubtedly play a crucial role in the star formation process. Nevertheless, the very existence of these disks, much less their physical properties, is not well-established. The importance of understanding these disks cannot be understated. For example, because of the larger masses involved, it is quite possible that high-mass disks are self-gravitating, unlike their Keplerian low-mass counterparts, implying that there may be significant dynamical differences between the high-mass and low-mass star formation processes.

S255IR is a part of the well known massive star-forming complex located between the evolved Sharpless H ii regions S255 and S257. Several authors have estimated the photometric distance to the complex at about 2.5 kpc (Russeil et al. 2007; Chavarría et al. 2008; Ojha et al. 2011). However, Rygl et al. (2010) report a distance of 1.6 kpc based on trigonometric parallax measurements of methanol masers. This seems to be the most accurate distance estimate and we adopt it here, too. The previous studies have shown that S255IR contains a cluster of early B-type stars (Howard et al. 1997; Itoh et al. 2001), several compact H ii regions (Snell & Bally 1986), and a number of H2 emission features (Miralles et al. 1997). The total mass of the S255IR core has been estimated from single-dish observations at M ∼ 300–400 M (Zinchenko et al. 2009; Wang et al. 2011). The estimate of the luminosity has been L ∼ 2 × 104 L (Wang et al. 2011).

The interferometric observations by Wang et al. (2011) revealed three continuum clumps and a high-velocity collimated outflow in the S255IR area. Our Paper I (Zinchenko et al. 2012) was devoted to the general structure and kinematics of the complex based on the Submillimeter Array (SMA) observations at 1.3 and 1.1 mm (the synthesized beam sizes were approximately 3farcs8 × 3farcs0 at 1.3 mm and 2farcs9 × 2farcs6 at 1.1 mm) as well as VLA ammonia observations and GMRT low frequency continuum data. Our results as well as data by Wang et al. (2011) indicate in particular a presence of a hot rotating core and a spectacular outflow in the S255IR region. However, the angular resolution was insufficient for a more detailed investigation of this core. In addition, a reliable evaluation of physical and chemical properties of this system required observations of additional molecular transitions. Therefore we performed observations of the S255IR area with the SMA at a much higher angular resolution and at higher frequencies.

Here we present observational data for S255IR obtained at 1.3 mm with the SMA in the very extended configuration and at 0.8 mm in the compact configuration. In addition we observed this area with the IRAM 30 m radio telescope in order to obtain short-spacing data complementing the SMA results. We mainly discuss properties of the dense cores and high velocity outflows observed in this area.

2. OBSERVATIONS AND DATA REDUCTION

2.1. SMA

The S255IR area was observed with the SMA in its compact configuration on 2010 December 14 at 350 GHz. A three-field mosaic was obtained with the following phase centers: 06h12m53fs800, 17°59'22farcs1, 06h12m54fs8876, 17°59'29farcs3353, and 06h12m52fs7124, 17°59'14farcs8647. The primary HPBW of the SMA antennas is 36'' at these frequencies. Typical system temperatures on source were between 200 and 400 K. The resulting uv coverage ranges from 10 to 90 kλ (8 antennas in the array). 3C454.3 and Uranus were used as the bandpass calibrators and 0532+075, 0750+125, and 0530+135 were used as the complex phase and amplitude gain calibrators. A total of 8 GHz (342−346 GHz in the LSB and 354−358 GHz in the USB) was observed with the SMA bandwidth doubling correlator configuration. The spectral resolution was 0.8125 MHz.

In addition on 2011 January 07 S255IR was observed in the very extended configuration at 225 GHz. A single field was observed with the same phase center as for the central field at 350 GHz. The primary HPBW of the SMA antennas is 55'' at these frequencies. Typical system temperatures on source were between 90 and 200 K. The resulting uv coverage ranges from 50 kλ to 350 kλ (7 antennas in the array). 3C454.3 and 3C279 were used as the bandpass calibrators and 0750+125 and 0530+135 were used as the complex phase and amplitude gain calibrators. A total of 8 GHz (217.0−220.8 GHz in the LSB and 229.1−232.95 GHz in the USB) was observed with the SMA bandwidth doubling correlator configuration. The spectral resolution was 0.406 and 1.625 MHz (given the limited correlator capability, we were not able to have a high resolution across the whole band, so we assigned different resolutions to different tracers). These data were combined with our previous data obtained at the same frequencies in the compact configuration (Paper I) which enables an investigation of the source structure in a wider range of spatial scales.

The gain calibrator flux scale, calibrated against Uranus at 350 GHz and Callisto at 225 GHz, was found to be consistent within 5% with the SMA calibrator database and estimated to be accurate within 20%.

The data calibration was carried out with the IDL superset MIR (Scoville et al. 1993), and subsequent imaging and analysis were done in MIRIAD (Sault et al. 1995). With robust weighting for the continuum and line data, the synthesized beam sizes are approximately 2farcs2 × 1farcs9 at 350 GHz and 0farcs5 × 0farcs4 at 225 GHz. The rms noise is approximately 7 mJy beam−1 and 1 mJy beam−1 in the continuum images at 350 GHz and 225 GHz, respectively, and 100 mJy beam−1 and 20 mJy beam−1 in the spectral cubes at these frequencies at 2 km s−1 resolution.

2.2. IRAM 30 m Radio Telescope

Single-dish observations of several molecular lines at the 30 m IRAM radio telescope were performed in 2012 October (N2H+ J = 3–2 at 279.5 GHz), 2012 December (SiO J = 5-4 at 217.1 GHz), and 2014 January (CO J = 3–2 at 345.8 GHz and CS J = 7–6 at 342.9 GHz). The antenna HPBW was 9farcs3, 11farcs9 and 7farcs5 at these frequencies, respectively. The observations were performed in the OTF mode with position switching using the HERA receiver at 217.1 GHz and EMIR receiver at higher frequencies. The central position of the maps was the same as for the primary field in the SMA observations. The reference position was selected at −500'' in right ascension from the central position. Apparently a weak CO(3–2) emission is present at the reference position resulting in a weak negative feature in the CO spectra at ${V}_{\mathrm{LSR}}\approx 24$ km s−1 (the bulk of the line emission in this area is observed at ${V}_{\mathrm{LSR}}$∼ 4–10 km s−1). This feature does not affect significantly the observed spectra. The system temperature was ∼200 K for the N2H+ observations, ∼300 K for the SiO observations and ∼600 K for the CO and CS observations. Pointing was checked regularly on nearby strong sources and pointing errors were within a few arcseconds. The antenna temperature calibration was made by the standard chopper-wheel method.

The map size was approximately 2farcm0 × 2farcm5 for the N2H+ and SiO observations, covering both S255IR and S255N regions, and approximately 1farcm5 × 1farcm0 for the CO and CS observations, covering only the S255IR area. Here we discuss only the data relevant to S255IR. The data on S255N are postponed to further publications.

The data reduction was performed with the GILDAS package (http://www.iram.fr/IRAMFR/GILDAS). Then the single-dish data were combined with the SMA data using the MIRIAD procedures as described by, e.g., Wang et al. (2011). The conversion to the flux density scale was made using the conversion factors presented on the IRAM 30 m telescope website.

3. OBSERVATIONAL RESULTS AND DATA ANALYSIS

With the SMA we detected several tens of spectral lines in both 350 and 225 GHz bands. A list of these lines, including their frequencies and energy of lower levels, is given in Tables 1, 2. The spectral line parameters are taken from the JPL (Pickett et al. 1998) and CDMS (Müller et al. 2001, 2005) catalogs.

Table 1.  List of Molecular Transitions Observed at the SMA in the Very Extended Configuration in S255IR

Molecule Transition Frequency El
    (GHz) (K)
12CO 2–1 230.538000 5.532
13CO 2–1 220.398684 5.289
CH3OH ${6}_{1}-{7}_{2}$ A 217.299202 363.496
  ${15}_{6}-{16}_{5}$ A 217.642677 735.160
  ${15}_{6}-{16}_{5}$ A+ 217.642678 735.160
  ${20}_{1}-{20}_{0}$ E 217.886504 497.919
  ${4}_{2}-{3}_{1}$ E 218.440063 34.976
  ${8}_{0}-{7}_{1}$ E 220.078561 86.051
  ${15}_{4}-{16}_{3}$ E 229.589056 363.420
  ${8}_{-1}-{7}_{0}$ E 229.758756 78.076
  ${19}_{5}-{20}_{4}$ A+ 229.864121 567.565
  ${19}_{5}-{20}_{4}$ A 229.939095 567.561
  ${3}_{-2}-{4}_{-1}$ E 230.027047 28.788
  ${22}_{2}-{21}_{-3}$ E 230.292196 598.499
  ${10}_{2}-{9}_{3}$ A 231.281110 154.248
  ${10}_{2}-{9}_{3}$ A+ 232.418521 154.248
  ${18}_{3}-{17}_{4}$ A+ 232.783446 435.360
13CS 5–4 231.220686 22.194
DCN 3–2 217.238530 10.425
HNCO ${10}_{\mathrm{0,10}}-{9}_{\mathrm{0,9}}$ 219.798282 47.471
  ${10}_{\mathrm{1,9}}-{9}_{\mathrm{1,8}}$ 220.584762 90.916
  ${10}_{\mathrm{1,10}}-{9}_{\mathrm{1,9}}$ 218.981170 90.569
  ${10}_{\mathrm{2,9}}-{9}_{\mathrm{2,8}}$ 219.733850 217.739
  ${10}_{\mathrm{2,8}}-{9}_{\mathrm{2,7}}$ 219.737193 217.739
  ${10}_{\mathrm{3,8}}-{9}_{\mathrm{3,7}}$ 219.656770 422.417
  ${10}_{\mathrm{3,7}}-{9}_{\mathrm{3,6}}$ 219.656771 422.417
HC3N 24–23 218.324711 120.504
SO ${6}_{5}-{5}_{4}$ 219.949433 24.429
H2CO ${3}_{\mathrm{0,3}}-{2}_{\mathrm{0,2}}$ 218.222195 10.483
  ${3}_{\mathrm{2,2}}-{2}_{\mathrm{2,1}}$ 218.475642 57.608
  ${3}_{\mathrm{2,1}}-{2}_{\mathrm{2,0}}$ 218.760071 57.613
OCS 18–17 218.903357 89.304
  19–18 231.060983 99.810
HCOOH ${10}_{\mathrm{0,10}}-{9}_{\mathrm{0,9}}$ 220.038072 48.061
  ${10}_{\mathrm{1,9}}-{9}_{\mathrm{1,8}}$ 231.505705 53.355
CH3CN ${12}_{6}-{11}_{6}$ 220.594423 315.313
  ${12}_{5}-{11}_{5}$ 220.641084 236.810
  ${12}_{4}-{11}_{4}$ 220.679287 172.555
  ${12}_{3}-{11}_{3}$ 220.709017 122.565
  ${12}_{2}-{11}_{2}$ 220.730261 86.849
  ${12}_{1}-{11}_{1}$ 220.743011 65.416
  ${12}_{0}-{11}_{0}$ 220.747261 58.272

Download table as:  ASCIITypeset image

Table 2.  List of Molecular Transitions Observed at the SMA in the Compact Configuration in S255IR

Molecule Transition Frequency El
    (GHz) (K)
12CO 3–2 345.795990 16.596
CS 7–6 342.882850 49.372
HCO+ 4–3 356.734242 25.682
CH3OH ${13}_{1}-{13}_{0}$ A ${}^{-+}$ 342.729796 211.024
  ${13}_{-1}-{14}_{-2}$ E 343.599019 607.550
  ${18}_{2}-{17}_{3}$ E 344.109039 402.884
  ${19}_{1}-{18}_{2}$ A+ 344.443433 434.697
  ${16}_{1}-{15}_{2}$ A 345.903916 316.049
  ${18}_{-3}-{17}_{-4}$ E 345.919260 442.829
  ${13}_{0}-{12}_{1}$ A+ 355.602945 193.959
HCN 4–3 354.505473 25.251
  4–3 (0,11c,0) 354.460435 1049.892
  4–3 (0,11d,0) 356.255568 1050.021
H13CN 4–3 345.339769 24.861
HC15N 4–3 344.200109 24.779
HC3N 38–37 345.609016 306.905
  39–38 354.697456 323.492
SO ${8}_{8}-{7}_{7}$ 344.310792 70.957
33SO ${8}_{\mathrm{9,8}}-{7}_{\mathrm{8,7}}$ 343.086102 61.564
  ${8}_{\mathrm{9,9}}-{7}_{\mathrm{8,8}}$ 343.087298 61.566
SO2 ${13}_{\mathrm{2,12}}-{12}_{\mathrm{1,11}}$ 345.338538 76.410
  ${12}_{\mathrm{4,8}}-{12}_{\mathrm{3,9}}$ 355.045517 93.960
  ${13}_{\mathrm{4,10}}-{13}_{\mathrm{3,11}}$ 357.165390 105.823
  ${15}_{\mathrm{4,12}}-{15}_{\mathrm{3,13}}$ 357.241193 132.537
  ${11}_{\mathrm{4,8}}-{11}_{\mathrm{3,9}}$ 357.387580 82.800
  ${8}_{\mathrm{4,4}}-{8}_{\mathrm{3,5}}$ 357.581449 55.202
  ${9}_{\mathrm{4,6}}-{9}_{\mathrm{3,7}}$ 357.671821 15.992
  ${7}_{\mathrm{4,4}}-{7}_{\mathrm{3,5}}$ 357.892442 47.835
  ${6}_{\mathrm{4,2}}-{6}_{\mathrm{3,3}}$ 357.925848 41.402
  ${17}_{\mathrm{4,14}}-{17}_{\mathrm{3,15}}$ 357.962905 162.932
H2CS ${10}_{\mathrm{0,10}}-{9}_{\mathrm{0,9}}$ 342.946424 74.132
  ${10}_{\mathrm{2,9}}-{9}_{\mathrm{2,8}}$ 343.322082 126.830
  ${10}_{\mathrm{3,8}}-{9}_{\mathrm{3,7}}$ 343.409963 192.613
  ${10}_{\mathrm{3,7}}-{9}_{\mathrm{3,6}}$ 343.414146 192.556
HCOOH ${16}_{\mathrm{1,16}}-{15}_{\mathrm{1,15}}$ 342.521225 127.153
  ${15}_{\mathrm{1,14}}-{14}_{\mathrm{1,13}}$ 343.952413 119.774
  ${16}_{\mathrm{0,16}}-{15}_{\mathrm{0,15}}$ 345.030596 126.494
  ${16}_{\mathrm{2,15}}-{15}_{\mathrm{2,14}}$ 356.137250 141.578

Download table as:  ASCIITypeset image

We present the results in the form of maps as well as spectra and line parameters at selected positions. For continuum observations we give positions, flux densities, and size estimates of the continuum sources.

3.1. Data Analysis

3.1.1. Methanol

For the methanol data analysis, we constructed simple radiative transfer model, which uses the large velocity gradient (LVG) approximation. Dust emission and absorption within the emission region was taken into account in the way described by Sutton et al. (2004). We assumed that the dust particles are intermixed with gas. The same physical temperature for the gas and dust components is assumed. The molecular emission region was assumed to be spherically symmetric and uniform in H2 density, gas and dust temperature, gas-to-dust ratio, and methanol fractional abundance. The influence of external infrared sources was not considered. The dust opacity law was chosen as ${\tau }_{\mathrm{dust}}\propto {\lambda }^{-2}$. We adopted a gas-to-dust mass ratio of 100 and a cross section at 1 mm of 2.6 × 10−25 cm2 (Sherwood et al. 1980). In addition to the model described in Sutton et al. (2004) we used the collision transition rates based on the model of collisions of methanol molecules with He and para-H2 molecules (Cragg et al. 2005). The scheme of energy levels in this model includes rotational levels with quantum numbers J up to 22, $| K| $ up to 9; the levels include the rotational levels of the ground, first, and second torsionally excited states. In total, 861 levels of A-methanol and 852 levels of E-methanol were considered according to Cragg et al. (2005).

With this model we made estimates of the hydrogen number density, ${n}_{{{\rm{H}}}_{2}}$, specific column density of methanol, ${N}_{{\mathrm{CH}}_{3}\mathrm{OH}}/{\rm{\Delta }}V$, gas kinetic temperature, Tkin, and fractional abundance of methanol ${N}_{{\mathrm{CH}}_{3}\mathrm{OH}}/{N}_{{{\rm{H}}}_{2}}$, using the measured values of "quasi-thermal" (i.e., non-maser) methanol lines. The variable parameter of the source size is introduced in order to take into account beam-dilution effects.

The brightness temperatures of all detected "quasi-thermal" methanol lines as well as upper limits for the brightness temperatures of other lines were taken into account.

We have searched for a set of parameters that provides the best agreement between the values of the calculated brightness temperatures (${T}_{i}^{\mathrm{mod}}$) and the measured brightness temperatures (${T}_{i}^{\mathrm{obs}}$). This corresponds to the minimum of ${\chi }^{2}\;={\displaystyle \sum }_{i}^{N}{(({T}_{i}^{\mathrm{obs}}-{T}_{i}^{\mathrm{mod}})/{\sigma }_{i})}^{2}$, where ${\sigma }_{i}$ is observational uncertainty for a particular line.

The sources appear inhomogeneous in some cases. However, the LVG approximation treats the problem locally and does not actually require physical homogeneity of the source. It can be well used for an inhomogeneous medium (e.g., Ossenkopf 1997). We have chosen the simplest approximation for the case of the source which is smaller than the beam—we introduced the filling factor which is equal to the portion of the source size in the beam. The value of the filling factor is well constrained because the model is sensitive to its changes. We do not have enough observational constraints to study real clumpiness of the source. Introducing additional poorly constrained parameters will reduce validity of the modeling.

3.1.2. Methyl Cyanide

Methyl cyanide (CH3CN) is a symmetric-top molecule which is an efficient "thermometer" for dense molecular gas (e.g., Boucher et al. 1980). To derive the kinetic temperature we used the population diagram analysis which takes into account the optical depth in the CH3CN lines and the beam filling factor as described by Wang et al. (2010).

3.2. Millimeter Wave Continuum

Maps of the continuum emission at 0.8 and 1.3 mm are presented in Figure 1. The 1.3 mm map was obtained by combining our new measurements in the very extended configuration with the previous measurements in the compact configuration (Paper I). It shows a rather extended emission and compact cores in the SMA1 and SMA2 clumps.

Figure 1.

Figure 1. Left panel: map of the 0.8 mm continuum emission in the S255IR. The contour levels are (1, 2, 3, 4, 5, 6, 7, 8, 9, 10) × 40 mJy beam−1. The dashed rectangle indicates the area shown in the right panel. The dotted circles indicate positions of two molecular clumps discussed in the text. Right panel: map of the 1.3 mm continuum emission in the S255IR. The contour levels are (2, 3, 4, 5, 7, 9, 11, 13) × 4 mJy beam−1. The SMA beam is shown in the lower left corner of both panels. The crosses mark the position of the SMA3 clump according to Wang et al. (2011).

Standard image High-resolution image

Estimates of the continuum source parameters are given in Tables 3, 4. For the measurements at 1.3 mm we indicate the parameters of the compact cores seen in the very extended configuration. Parameters of the more extended components were presented in Paper I. The measured fluxes are several hundreds mJy at 350 GHz and several tens mJy at 225 GHz in the very extended configuration. The deconvolved sizes are from ∼1 arcsec to a few arcseconds at 345 GHz and about 0.3 arcsec for the compact cores at 225 GHz.

Table 3.  Names, Positions, Flux Densities, Deconvolved Angular Sizes, and Position Angles of the Millimeter Wave Continuum Sources Measured at 350 GHz in the Compact Configuration

Name α(2000) δ(2000) ${S}_{350}$ ${\theta }_{\mathrm{max}}$ ${\theta }_{\mathrm{min}}$ P.A.
  (h m s) (${}^{\circ }$ ${}^{\prime }$ ${}^{\prime\prime }$) (Jy) (${}^{\prime\prime }$) (${}^{\prime\prime }$) (${}^{\circ }$)
S255IR-SMA1 6:12:54.00 17:59:23.2 0.50 1.4 0.6 −11
S255IR-SMA2 6:12:53.76 17:59:26.1 0.56 2.4 2.1 4
S255IR-SMA3 6:12:53.86 17:59:23.7 0.13 1.3 0.3 44
S255IR-SMA4 6:12:54.01 17:59:12.0 0.25 5.9 3.8 −82

Download table as:  ASCIITypeset image

Table 4.  Names, Positions, Flux Densities, Deconvolved Angular Sizes, and Position Angles of the Millimeter Wave Continuum Sources Measured at 225 GHz in the Very Extended Configuration

Name α(2000) δ(2000) ${S}_{225}$ ${\theta }_{\mathrm{max}}$ ${\theta }_{\mathrm{min}}$ P.A.
  (h m s) (${}^{\circ }$ ${}^{\prime }$ ${}^{\prime\prime }$) (Jy) (${}^{\prime\prime }$) (${}^{\prime\prime }$) (${}^{\circ }$)
S255IR-SMA1 6:12:54.010 17:59:23.06 0.058 0.30 0.27 −19
S255IR-SMA2 6:12:53.779 17:59:26.16 0.016 0.49 0.31 −27

Download table as:  ASCIITypeset image

3.3. Basic Properties of the Molecular Emission

Here we give the general description of the observed molecular emission. More detailed information on the relevant species and transitions is presented in the following sections.

The general morphology and kinematics of the molecular emission were described in Paper I. Our new data set contains several tens of molecular transitions. In Figure 2 we present representative maps obtained with the SMA at 0.8 mm.

Figure 2.

Figure 2. Maps of the integrated line emission for several representative transitions of various molecules (contours) overlaid on the image of the 0.8 mm continuum emission. The velocity interval is shown at the bottom of each panel. The dashed contours show negative features due to the missing flux. The contour levels are the following: (−3, −2, 2, 3, 5, 7, 9, 12, 15, 20) × A Jy beam−1 km s−1, where A = 3 for HCN and HCO+, A = 1 for H13CN, HC15N, HC3N, H2CS, CH3OH, 33SO, SO2, and HCOOH, A = 2 for CS, and A = 1.5 for SO. The SMA beam is shown in the lower left corner of each panel.

Standard image High-resolution image

These maps confirm that SMA1 is the brightest source of molecular emission in this area. High-excitation lines are observed exclusively toward this clump. Several lower excitation molecular lines are detected also in SMA2. Emission of HCN, SO, and CH3OH is also observed in SMA4. HCN and CS emission is probably present in the area to the west of SMA2, designated as S255IR-N2H+(1) in Paper I. Since it is observed not in N2H+ only and for simplicity we shall designate it hereafter as SMA2-W.

A new feature, not noticed in our Paper I and in other previous works is a molecular clump to the east of SMA1 with a rather strong emission in the HCN, HCO+, CS and SO lines (Figure 2). We shall designate it as SMA1-E. This clump is located near the head of the jet observed in the NIR emission (see below).

Spectra of several representative transitions toward the SMA1 and SMA2 clumps are presented in Figures 3, 4.

Figure 3.

Figure 3. Spectra of several representative molecular transitions toward the SMA1 clump. The CO(3–2) spectrum is scaled by a factor of 0.5. It is obtained from the combined SMA and IRAM 30 m data. The other spectra are from the SMA data only. The angular resolution is about 2 arcsec.

Standard image High-resolution image
Figure 4.

Figure 4. Spectra of several representative molecular transitions toward the SMA2 clump. The CO(3–2) spectrum is scaled by a factor of 0.5. It is obtained from the combined SMA and IRAM 30 m data. The other spectra are from the SMA data only. The angular resolution is about 2 arcsec.

Standard image High-resolution image

As mentioned in Paper I the main emission peak toward the SMA1 clump is at ${V}_{\mathrm{LSR}}$ ∼ 4–5 km s−1. The line width is $\gtrsim 5$ km s−1. Several lines including CO, HCN, HCO+, CS, and SO show broad wings indicative of high-velocity outflow. We discuss this feature in the following sections. A rather strong emission is detected from vibrationally excited HCN with an excitation energy ∼1000 K above the ground level.

The line emission from the SMA1-E clump peaks at about 8 km s−1. The line width is large, ∼5 km s−1.

The emission from the SMA2 clump is observed at about 10 km s−1. The lines are narrow, ∼2 km s−1.

With the very extended array we detected line emission almost exclusively from the SMA1 compact core. No emission was detected in C18O and SiO indicating an absence of compact structures in the lines of these molecules. The results of these observations are presented and discussed below.

4. STRUCTURE, KINEMATICS, AND PHYSICAL PROPERTIES OF DENSE CORES

Our data presented in Paper I and here as well as data by Wang et al. (2011) indicate the presence of four continuum clumps in the S255IR area, designated S255IR-SMA1 to S255IR-SMA4. In addition, in Paper I we detected clumps with a rather strong molecular emission without continuum counterparts in the SMA data. One of them, S255IR-N2H+(1), is located close to SMA2 and was observed in the N2H+, NH3, and several CH3OH lines. Estimates of their basic physical properties were presented in Paper I. Here we further investigate these objects using the new data set.

4.1. SMA1

SMA1 is the brightest object in this area. As shown in Section 3 most of the observed molecular lines are detected only here. The deconvolved size of the continuum source at 350 GHz measured with the SMA in the compact configuration (Table 3) is close to that found in Paper I. The flux density measured at 350 GHz is only slightly higher than the flux density obtained at 284 GHz (Paper I). It is worth noting that in Paper I we could not separate the SMA1 and SMA3 clumps. However, even if we take the integrated flux of these clumps at 350 GHz, the spectral index in the range 284–350 GHz will be only 1.6 which is inconsistent with a presumably optically thin dust emission at these frequencies. Most probably this implies a significantly larger flux loss at 350 GHz in comparison with the measurements at 284 GHz due to a smaller beam size.

The deconvolved size of the continuum source detected with the SMA in the very extended configuration is about 0farcs3 (Table 4) which corresponds to about 500 AU. The observed aspect ratio for this core is close to unity.

4.1.1. Kinematics

Wang et al. (2011) noticed rotation of the core around the axis roughly parallel to the outflow direction. Now, at sub-arcsecond resolution we can investigate the core kinematics on smaller scales. In Figure 5 we present maps of the first moment of the CH3OH emission in the ${4}_{2}-{3}_{1}$ E line and CH3CN emission in the ${12}_{3}-{11}_{3}$ line. In this figure we also indicate the axis of the jet previously identified through IR observations (Howard et al. 1997) and locations of the water masers measured with the VLBA (Goddi et al. 2007).

Figure 5.

Figure 5. Maps of the first moment of the CH3OH emission in the ${4}_{2}-{3}_{1}$ E line (left panel, color scale) and CH3CN emission in the ${12}_{3}-{11}_{3}$ line (right panel, color scale) toward the SMA1 clump. The intensity cut-off is 100 mJy beam−1 km s−1 for CH3OH and 200 mJy beam−1 km s−1 for CH3CN. Contours show the 1.3 mm continuum emission. The triangles mark water masers (Goddi et al. 2007) and the arrows indicate their proper motions (in the cases when they are measured). The red line indicates the jet axis as found by Howard et al. (1997). The beam for the molecular maps is shown in the lower left corner of both panels. The scale for the velocities of the proper motions is shown in the lower right corner. The pixels with the lowest velocities appear as the blue squares in the right panel.

Standard image High-resolution image

This figure clearly shows that the core is really rotating around the axis of the jet. The rotation velocities along the line of sight amount to a few km s−1 (it varies from about 3 km s−1 to about 7 km s−1 for CH3OH ${4}_{2}-{3}_{1}$ E and to about 8 km s−1 for CH3CN ${12}_{3}-{11}_{3}$). In Figure 6 we present the position–velocity diagram for the CH3OH emission in the ${4}_{2}-{3}_{1}$ E line along the cut through the core center perpendicular to the jet axis (P.A. = 157°). It clearly shows the velocity gradient which can be consistent with Keplerian rotation but the angular resolution is still insufficient for firm conclusions about the character of this rotation.

Figure 6.

Figure 6. Position–velocity diagram for the CH3OH emission in the ${4}_{2}-{3}_{1}$ E line along the cut through the SMA1 center perpendicular to the jet axis (P.A. = 157°). The dashed curves correspond to Keplerian rotation around a central mass of 20 M with an inclination angle of 25°, and the dotted curves correspond to the inclination angles of 25° ± 5°.

Standard image High-resolution image

An implicit indication of a probable further increase of the rotation velocity in the innermost part of the core comes from the line width map (Figure 7). It shows a significant increase of the line width in the center which can be related to rotation (the line width increases from about 2 km s−1 at the periphery to about 6 km s−1 in the center).

Figure 7.

Figure 7. Map of the second moment of the CH3OH emission in the ${4}_{2}-{3}_{1}$ E line (color scale). Contours show the 1.3 mm continuum emission. The beam for the molecular map is shown in the lower left corner.

Standard image High-resolution image

It is also worth mentioning an increased line width along the jet axis (Figure 7). It can be probably explained by an increased turbulence caused by the passage of the jet.

The HCO+(4–3) and probably HCN(4–3) spectra toward SMA1 (Figure 3) show the redshifted absorption dip which may be suggestive of infall. The CO(3–2) spectrum measured with the SMA has a similar (although broader) feature (Figure 8). At the same time the CO(3–2) spectrum obtained from the combined SMA and IRAM 30 m data shows a peak at these velocities. Most probably it means that this feature is related to an extended component resolved out by the SMA, although an infall cannot be fully excluded.

Figure 8.

Figure 8. SMA spectra of HCO+(4–3) and CO(3–2) as well as the combined SMA+30 m spectrum of CO(3–2) toward SMA1. The angular resolution is about 2 arcsec.

Standard image High-resolution image

4.1.2. Physical Properties

Let us consider physical parameters of this core. We estimated gas kinetic temperature from the CH3CN and CH3OH observations. Modeling of the CH3CN emission (Figure 9) as described in Section 3.1.2 yields a kinetic temperature in the range (74.9–197.6) K (1σ confidence interval) toward the emission peak (Table 5). We have detected a large number (15) of CH3OH transitions toward SMA1 (Table 1). From the CH3OH analysis (Section 3.1.1) we obtain ${T}_{\mathrm{kin}}=178$ K with the 165–195 K 1σ confidence interval at the center, ${T}_{\mathrm{kin}}=170-200$ K at 0farcs2 to the north, and ${T}_{\mathrm{kin}}=140-165$ K at 0farcs2 to the south (Table 6). Therefore there is probably a temperature gradient in the north–south direction. The temperature seems to be anti-correlated with the CH3OH column density (Table 6, Figure 10).

Figure 9.

Figure 9. CH3CN spectrum measured toward the SMA1 emission peak and the model fit.

Standard image High-resolution image
Figure 10.

Figure 10. Example of the CH3OH data fitting for the central position in SMA1. Variations of χ2 with the model parameters are shown in the two top and left bottom panels. The minimum of χ2 is marked by the star. Confidence intervals of 0.25, 1, 1.6, and 2.6 σ are plotted on the gray scale. The confidence interval of 1 σ is marked by the white line. A comparison of the model (stars) and observed main-beam temperatures is shown in the right bottom panel.

Standard image High-resolution image

Table 5.  Kinetic Temperature, CH3CN Column Density, Beam Filling Factor, and Line Optical Depth Derived from the CH3CN Data Obtained in the Very Extended Configuration (HPBW $\approx 0\buildrel{\prime\prime}\over{.} 4$) at the Central Position in SMA1

Parameter Fit results
Tkin (K) 128.8 (74.9–197.6)
${N}_{{\rm{L}}}({\mathrm{CH}}_{3}\mathrm{CN})$ (1016 cm−2) 2.12 (0.96–5.80)
Beam filling factor 0.23 (0.18–0.36)
$\tau (K=0)$ 2.53 (0.93–8.37)

Note. The 1σ confidence intervals are indicated in parentheses.

Download table as:  ASCIITypeset image

Table 6.  Physical Parameters Derived from the CH3OH Data Obtained in the Very Extended Configuration (HPBW $\approx 0\buildrel{\prime\prime}\over{.} 4$) at Several Positions in SMA1

Offsets Tkin ${N}_{{\mathrm{CH}}_{3}\mathrm{OH}}/{\rm{\Delta }}V$ ${n}_{{{\rm{H}}}_{2}}$ Beam Filling Factor ${N}_{{\mathrm{CH}}_{3}\mathrm{OH}}/{N}_{{{\rm{H}}}_{2}}$
(${}^{\prime\prime }$, ${}^{\prime\prime }$) (K) (1012 cm−3s) (108 cm−3) (%)  
0, +0.2 183 (170–200) 3.6 (2.8–5.0) 3.2 14.8 (13.9–16.4) ${10}^{-6}(\gt {10}^{-7})$
0, 0 178 (165–195) 5.6 (4.0–9.5) 3.2 16.0 (15.2–17.2) ${10}^{-6}(\gt {10}^{-7})$
0, −0.2 153 (140–165) 8.9 (5.0–12.6) 0.2 15.2 (13.3–16.0) ${10}^{-6}(\gt {10}^{-7})$

Note. The 1σ confidence intervals are indicated in the parentheses (the density ${n}_{{{\rm{H}}}_{2}}$ is not constrained by the model as evident from Figure 10).

Download table as:  ASCIITypeset image

We adopt the gas kinetic temperature of 170 K as the average value (it is close to the weighted average of the CH3OH and CH3CN results). Assuming the same temperature (170 K) for the dust we obtain a total mass of this hot component of about 0.3 M. As in Paper I we assume a gas-to-dust mass ratio of 100 and adopt a dust absorption efficiency following Ossenkopf & Henning (1994). The peak gas column density estimated from the continuum data is $N({{\rm{H}}}_{2})\sim 3\times {10}^{24}$ cm−2. With a size of 500 AU the mean density of hot gas is about 6 × 108 cm−3. This is two times higher than our estimate of the SMA1 mean density in Paper I taking into account the difference in the adopted distances here and in Paper I. This increase can be expected since now we consider much smaller scales near the core center. The CH3OH modeling puts no significant constraints on density. The relative abundance of CH3OH is ∼10−6.

Both CH3CN and CH3OH observations indicate a beam filling factor of 0.15–0.2. This means that the source is very inhomogeneous on the 0farcs4 scale and probably consists of clumps with a size of ≪500 AU and a density much higher than the mean density found above. This higher density estimate does not contradict the CH3OH data. A similar picture of clumpy medium was inferred from our observations of high-mass star-forming regions on larger scales (Pirogov & Zinchenko 2008; Pirogov et al. 2012). Most probably it reflects turbulence in the cores.

Our data set includes several other tracers of the hot gas. One of the most important is HNCO. We detected HNCO lines in different ${K}_{-1}$ ladders (${K}_{-1}=0,1,2,3$) with excitation energies up to ∼400 K (Table 1). In Figure 11 we present the rotational diagram for the observed HNCO transitions. This diagram was obtained from peak integrated line intensities found by 2D Gaussian fitting of the integrated intensity maps in different HNCO lines. We used the line strengths and dipole moment from the Cologne Database for Molecular Spectroscopy (Müller et al. 2001, 2005).

Figure 11.

Figure 11. Rotational diagram for HNCO toward the emission peak.

Standard image High-resolution image

This diagram indicates a rotational temperature of 318 ± 70 K (Table 7). However, the ${K}_{-1}=0,1$ transitions can be saturated. The peak brightness temperature in these transitions is close to the brightness temperature in the apparently optically thick lines (e.g., CO and CH3CN) and our modeling using RADEX (van der Tak et al. 2007) shows that at the derived physical parameters and column density the optical depth in these lines is about unity. In this case the derived rotational temperature would apparently represent an upper limit to the excitation temperature. The deconvolved source size is ∼0farcs3 × 0farcs2 for the ${K}_{-1}=0$ transition and decreases for higher ${K}_{-1}$ ladders. The map of the integrated intensity in the HNCO ${K}_{-1}=2$ transition is shown in Figure 12 along with the map of the OCS emission which is another tracer of hot gas. The peaks of the HNCO and OCS emission practically coincide with the continuum peak, although distributions of these molecules seem to be somewhat different.

Figure 12.

Figure 12. Maps of the integrated intensity of HNCO ${J}_{{{\rm{K}}}_{-1}}={10}_{2}-{9}_{2}$ (left panel) and OCS $J=18-17$ (right panel) transitions (color scale) toward the SMA1 clump. The intensity units are Jy beam−1 km s−1. The beam size is 0farcs4. The contours show the 1.3 mm continuum emission with the same angular resolution. The beam for the molecular maps is shown in the lower left corner of both panels.

Standard image High-resolution image

Table 7.  Fit Results for the Rotational Diagrams Obtained toward SMA1

Molecule Tkin $\mathrm{log}{N}_{{\rm{L}}}$ HPBW
  (K) (cm−2) (${}^{\prime\prime }$)
HNCO 318(70) 16.34(0.04) $\approx 0.4$
SO2 146(16) 15.50(0.05) $\approx 2$
  65(11) 15.35(0.08) $\approx 2$

Note. The 1σ uncertainties are indicated in parentheses. In the last column the approximate HPBW for the corresponding data is indicated.

Download table as:  ASCIITypeset image

Table 8.  Physical Parameters Derived from the CH3OH Data Obtained in the Compact Configuration (HPBW $\approx 3^{\prime\prime} $)

Object Tkin ${N}_{{\mathrm{CH}}_{3}\mathrm{OH}}/{\rm{\Delta }}V$ ${n}_{{{\rm{H}}}_{2}}$ Beam Filling Factor ${N}_{{\mathrm{CH}}_{3}\mathrm{OH}}/{N}_{{{\rm{H}}}_{2}}$
  (K) (1012 cm−3s) (cm−3) (%)  
SMA2 50 (40–80) 2.0 (1.8–3.5) $6.3\times {10}^{6}$ ($\leqslant {10}^{8}$) 99.3 ($\geqslant 84$) ${10}^{-8}(\geqslant {10}^{-8})$
SMA2-W 40 (25–65) 3.2 (2.2–4.5) $1.0\times {10}^{6}$ ($6.3\times {10}^{4}$$2.5\times {10}^{6}$) 64.0 (63–70) ${10}^{-7}(\geqslant {10}^{-8})$
SMA4 45 (30–90) 3.6 (0.8–7.9) $5.6\times {10}^{4}$ (103$6.3\times {10}^{5}$) 99.9 ($\geqslant 20$) ${10}^{-7}({10}^{-6}$${10}^{-9})$

Note. The 1σ confidence intervals are indicated in parentheses.

Download table as:  ASCIITypeset image

The total HNCO column density derived from the rotational diagram is ∼2 × 1016 cm−2. We used the partition function from the Cologne Database for Molecular Spectroscopy (Müller et al. 2001, 2005), too. A comparison with the total gas column density indicates HNCO abundance of $X(\mathrm{HNCO})\sim {10}^{-8}$. This is a rather high value, close to the highest HNCO abundance derived in the survey of massive cores by Zinchenko et al. (2000). However, taking into account the note above about the HNCO lines saturation, this value may need a correction.

In the 350 GHz band we detected many SO2 lines (Table 2). The corresponding population diagram is shown in Figure 13. It indicates a range of temperatures. For low-excitation transitions the rotational temperature is 65 ± 11 K. For the high-excitation ones it is 146 ± 16 K. The SO2 observations were performed at a much lower angular resolution (2 arcsec) than in the case of HNCO and apparently include emission from both the hot core and the surrounding cloud. The derived temperatures are consistent with the other estimates for these components.

Figure 13.

Figure 13. Population diagram for SO2 at the SMA1 emission peak.

Standard image High-resolution image

Another indicator of the hot environment is the vibrationally excited HCN. We detected the emission in the ${v}_{2}=1$ state, about 1000 K above the ground state. For the ${v}_{2}=2$ emission the upper limit is about five times lower. Following the analysis presented in, e.g., Veach et al. (2013) we obtain an upper limit for the excitation temperature between these states of about 500 K. This is consistent with the other estimates of the gas temperature given above. The critical density for HCN excitation is high (>1010 cm−3, Veach et al. 2013) which is consistent with our density estimates for the hot gas.

4.1.3. A Cold Clump in the Hot Core?

In Figure 14 we plot maps of the DCN $J=3-2$ and 13CS $J=5-4$ integrated line emission in the SMA1 core. The emission regions are very compact and the emission peak is shifted from the continuum peak which apparently coincides with the YSO location. The projected distance from the continuum peak is roughly 300 AU.

Figure 14.

Figure 14. Maps of the integrated intensity of the DCN $J=3-2$ (left panel) and 13CS $J=5-4$ (right panel) transitions (color scale) toward the SMA1 clump. The intensity units are Jy beam−1 km s−1. The beam size is 0farcs4. The contours show the 1.3 mm continuum emission with the same angular resolution. The beam for the molecular maps is shown in the lower left corner of both panels.

Standard image High-resolution image

The velocities of the DCN and 13CS emission are practically the same as those of high excitation lines of other molecules tracing apparently the hot gas. Therefore, most probably the observed DCN emission arises within the disk and since we probably see the disk nearly face-on as discussed above, the physical distance from the center is not much larger than the projected distance. An estimate of the abundances in the LTE approximation gives $X(\mathrm{DCN})\gt {10}^{-11}$ and $X{(}^{13}\mathrm{CS})\gt 3\times {10}^{-11}$ (assuming the total mass of the clump <10 M and temperature of 50–100 K). This is a "normal" value for 13CS while the derived DCN abundance implies a significant deuteriation as follows from comparison with typical HCN abundances in massive cores (e.g., Zinchenko et al. 2009). As pointed out in Paper I, recent modeling by Albertsson et al. (2013) shows that the DCN/HCN abundance ratio sharply drops at temperatures ≳80 K. It means that the temperature of the DCN-emitting clump should be rather low, much lower than the temperature of the hot gas in the disk. Another possible explanation, as also discussed in Paper I, could be a very young age of the clump, insufficient to reach the steady-state DCN/HCN abundance ratio, but this looks less probable. In principle, interiors of the accretion disk can be well shielded from an external radiation, probably providing necessary low temperatures. In any case this clump deserves further investigation in order to clarify its properties and nature. It is worth mentioning that the mass of this clump cannot be lower than ∼1 M, otherwise the inferred 13CS abundance would be unrealistically high. The virial mass we estimate about the same, hence the clump can be gravitationally bound.

4.2. SMA2

As in case of the SMA1 the spectral slope for the continuum emission in the frequency range 284–350 GHz is too low (∼1.3), which is probably caused by a higher flux loss at higher frequencies. With the SMA at sub-arcsecond resolution we detected a compact component in the continuum emission of about the same size as in the SMA1. However, no high-excitation molecular lines could be detected in this area. In Paper I we derived the temperature of the SMA2 clump ∼40 K. Now with the extended data set, from the methanol excitation analysis (Section 3.1.1) we found the 1σ confidence interval for temperatures of 40–80 K (Table 8). The best estimate of the methanol relative abundance is $X({\mathrm{CH}}_{3}\mathrm{OH})\sim {10}^{-8}$. These estimates are obtained at the 2 arcsec scale. It cannot be excluded that the temperature of the compact structure observed in the very extended configuration is somewhat higher. However, it cannot be much higher since there is no sign of a higher temperature component in the molecular data. Assuming a temperature of 50 K we obtain the mass of the compact core ∼0.2 M.

To the west from the SMA2 there is an area of molecular emission without a continuum counterpart in the SMA data, observed earlier in the N2H+, NH3, and several CH3OH lines (Paper I and Wang et al. 2011). Our new methanol data analysis indicates a temperature in the range 25–65 K and a density $n({{\rm{H}}}_{2})$ in the range 6.3 × 104–2.5 × 106, $X({\mathrm{CH}}_{3}\mathrm{OH})\sim {10}^{-7}$ (Table 8). Some of the methanol transitions are inverted in the model and can be masing. These high gas densities and rather high methanol abundances at relatively low temperatures can be explained by the influence of shock.

4.3. SMA3

In Paper I we did not detect this component which coincides with the near-infrared source NIRS 1 (Tamura et al. 1991) and is identified as a massive disk candidate by NIR polarization observations (Jiang et al. 2008). However, Simpson et al. (2009) argued that these measurements could be affected by instrumental effects and there is no real evidence for the "polarization disk" here. At the same time the NIR measurements indicate a possible outflow related to this object. A comparison with the measurements by Wang et al. (2011) gives a spectral index of about 3.3 in the frequency range 225–350 GHz, consistent with the optically thin dust emission. Wang et al. (2011) estimated the mass of this clump at about 2 M. At sub-arcsecond resolution we do not see any continuum (Figure 1) or molecular emission at this position.

4.4. SMA4

This clump was first detected in Paper I. It shows a weak continuum and spectral line emission in several molecular transitions. The measured continuum flux density at 350 GHz is practically the same as at 284 GHz which also implies a much higher flux loss at the higher frequency. The temperature derived from our new methanol data analysis is about 45 K (Table 8), higher than estimated in Paper I from ammonia data. However, the ammonia emission here is very weak and the uncertainties are high. Some methanol transitions are inverted in the model. The influence of shock is also probable.

5. MORPHOLOGY AND PROPERTIES OF THE OUTFLOWS

In the S255IR area high velocity emission is detected in lines of CO, SiO and several high density tracers including HCN, HCO+, and CS. In Figures 15, 16 we present channel maps of the CO $J=3-2$ emission in the blue and red line wings, respectively. The maps of the integrated line wing emission (Figure 17) show that the CO high velocity emission observed with the SMA looks like a highly collimated bipolar outflow originating near SMA2. There is another more compact component near SMA1.

Figure 15.

Figure 15. Channel maps of the CO blue wing emission in the S255IR. The numbers in the upper left corner indicate the channel velocity in km s−1. The dashed contours show negative features due to the missing flux. The contour levels are (−3, 3, 5, 7, 10, 15, 20, 30, 40) × 100 mJy beam−1. The SMA beam is shown in the lower left corner of each panel.

Standard image High-resolution image
Figure 16.

Figure 16. Channel maps of the CO red wing emission in the S255IR. The numbers in the upper left corner indicate the channel velocity in km s−1. The dashed contours show negative features due to the missing flux. The contour levels are (−3, 3, 5, 7, 10, 15, 20, 30, 40) × 100 mJy beam−1. The SMA beam is shown in the lower left corner of each panel.

Standard image High-resolution image
Figure 17.

Figure 17. Left panel: maps of the CO(3–2) high velocity emission as observed with the SMA (blue and red contours) in the S255IR area overlaid on the continuum image at 0.8 mm. The SMA1 and SMA2 continuum clumps are marked. The velocity intervals are from −63 to −13 km s−1 for the blue wing and from 25 to 75 km s−1 for the red wing. The dashed rectangle indicates the area shown in the right panel. Right panel: maps of the HCO+(4–3) high velocity emission (blue and red contours) in the S255IR area overlaid on the continuum image at 0.8 mm. The velocity intervals are from −19 to 3 km s−1 for the blue wing and from 11 to 27 km s−1 for the red wing. The black contours show the Fe ii emission (Wang et al. 2011). The map of the 15 GHz continuum emission (from the VLA archival data, the angular resolution is about 4 arcsec) is plotted with green contours. The dashed line indicates the jet axis (P.A .= 67°) as in Figure 5. The SMA beam is shown in the lower left corner of both panels.

Standard image High-resolution image

On the other hand the bipolar outflow observed in HCN and HCO+ is apparently associated with SMA1 while is parallel to the CO flow (Figure 17). It is apparently associated with the jet observed in particular in the Fe ii emission which is also plotted in this figure. It is worth mentioning that the extent of the jet is much smaller than the extent of the outflow seen in CO, although their orientations coincide.

The SMA data alone hint at two parallel outflows with different origins. However, single-dish CO(3–2) observations with the IRAM 30 m telescope show a different picture. Here we see a wider outflow clearly originating at SMA1 (Figure 18). While the emission peaks coincide in the SMA and 30 m maps, near the driving source the CO emission measured with the SMA traces only the northern edge of the outflow. This shows that the SMA map gives a distorted picture and in reality we have here a wider angle outflow originating at SMA1.

Figure 18.

Figure 18. Maps of the CO(3–2) high velocity emission as observed with the IRAM 30 m telescope (blue and red thick contours) in the S255IR area overlaid on the continuum image at 0.8 mm. The thin contours show the SMA maps from Figure 17. The dashed line indicates the axis of the position–velocity cut (P.A. = 67°). The orange contours show the N2H+(3–2) integrated line emission obtained by combining the SMA and 30 m data. The cross marks the position of the high-velocity dense clump (Section 5.1). The 30 m beam and the SMA beams for the CO and N2H+ observations, respectively, are shown in the lower left corner (from left to right).

Standard image High-resolution image

Apparently this means that this northern edge of the outflow contains a relatively bright component with a characteristic scale comparable to the SMA beam. The southern edge of the outflow near the driving source would be more diffuse and resolved out by the SMA. Most probably this implies that the CO emission is formed in a compressed layer surrounding the outflow cavity. For some (unclear) reason in the northern part this layer is more pronounced. It is worth mentioning that the brightest peaks of the high velocity CO emission practically coincide in the SMA and 30 m maps and also coincide with spots of molecular hydrogen emission (Wang et al. 2011) which probably indicate bow shocks. Basic physical parameters of the outflow were estimated by Wang et al. (2011) from their CO(2–1) data. Our new data are consistent with these estimates.

The position–velocity (P–V) diagram for the outflow along the cut indicated in Figure 18, constructed from the 30 m data, is shown in the left panel of Figure 19. The gap at ${V}_{\mathrm{LSR}}\sim 24$ km s−1 is caused by the emission at the reference offset position mentioned above. In the right panel we present the P–V diagram obtained from the SMA CO(3–2) data. The main features of these diagrams coincide. They show non-monotonic dependence of the velocity on offset, which hints at two outflow components at different distances from the driving source. These P–V diagrams are somewhat different from that presented by Wang et al. (2011). However, Wang et al. (2011) plotted this diagram for a different cut, with the position angle of 75°, which makes the direct comparison irrelevant.

Figure 19.

Figure 19. Left panel: the position–velocity map for the CO(3–2) emission as observed with the IRAM 30 m telescope along the cut indicated in Figure 18. The contour levels are on a logarithmic scale from 0.4 to 31.5 K. The gap at ${V}_{\mathrm{LSR}}\sim 24$ km s−1 is caused by the emission at the reference offset position mentioned above. Right panel: the position–velocity map for the CO(3–2) emission as observed with the SMA along the same cut. The contour levels are on a logarithmic scale from 0.5 to 11.5 Jy beam−1.

Standard image High-resolution image

In Figure 20 we plot the P–V diagram for the HCO+(4–3) emission as observed with the SMA, along with the part of the CO(3–2) P–V diagram for the same intervals of the offset and velocity. The diagrams are very similar, which means that both molecules trace apparently the same gas.

Figure 20.

Figure 20. Left panel: the position–velocity map for the HCO+(4–3) emission as observed with the SMA along the cut indicated in Figure 17. The contour levels are on a logarithmic scale from 0.4 to 7.6 Jy beam−1. Right panel: a part of the position–velocity map for the CO(3–2) emission measured by the SMA from Figure 19 for the same intervals of the offset and velocity as in the left panel.

Standard image High-resolution image

5.1. Dense High Velocity Clump

Our data show a strong, compact blue-shifted CS and HCN emission close to the peak of the CO blue-shifted line wing emission. The peak of this CS and HCN emission is marked by cross in Figure 18. The CO(3–2), CO(2–1), HCN(4–3), CS(7–6), HCO+(4–3), and N2H+(3–2) spectra toward this position are presented in Figure 21.

Figure 21.

Figure 21. CO(3–2), CO(2–1), HCN(4–3), CS(7–6), HCO+(4–3), and N2H+(3–2) spectra toward the position marked by cross in Figure 18.

Standard image High-resolution image

One can see rather strong and broad HCN and CS lines at a central velocity of about −5 km s−1. There is a hint of wings in these lines extending to about −20 km s−1. The peak of the CO emission, especially in the higher $J=3-2$ transition is observed at more negative velocities. There is no detectable HCO+(4–3), N2H+(3–2), CH3CN, and continuum emission (for N2H+ and CH3CN we use the data from Paper I).

The spatial distribution of the CS(7–6) and CO(3–2) emission integrated in the velocity ranges −13...+1 km s−1 and −19...+1 km s−1, respectively, is shown in Figure 22. The HCN distribution is very similar to that of CS. The deconvolved size of the CS emitting clump is about 1farcs1 × 0farcs3 which corresponds to 1800 AU × 500 AU. It is apparently located at the head of the stream observed in CO, almost exactly at the jet axis shown in Figure 18.

Figure 22.

Figure 22. Map of CO(3–2) emission in the velocity range −19...+1 km s−1 (contours) overlaid on an image of CS(7–6) emission in the velocity range −13...+1 km s−1 (gray scale). The intensity units for the image are Jy beam−1 km s−1. The contour levels are (3, 5, 7, 9, 11, 13) × 5 Jy beam−1 km s−1. The SMA beam is shown in the lower left corner.

Standard image High-resolution image

At the velocity of the CS and HCN emission peak, the brightness temperatures in the CO(3–2), CO(2–1), and HCN(4–3) lines are practically the same. The frequencies of the CO(3–2), HCN(4–3), and CS(7–6) lines are close to each other and beam parameters for them should be similar, too. However, the comparison with the CO(2–1) is complicated by the significantly different frequencies.

The CO(3–2) line is most probably saturated. Then the HCN(4–3) line should be saturated, too. The CS line is weaker implying either a relatively lower optical depth (but close to unity anyway) or a smaller size of the emission region. A simple modeling using, e.g., RADEX (van der Tak et al. 2007) shows that the column densities of HCN and CS required to explain such optical depths are ≳1015 cm−2. The relative abundances of these molecules are ≤10−8. Therefore the total gas column density in the clump is ≳1023 cm−2. Then, using the size estimated above we obtain the gas density n ≳ 3× 106 cm−3. This estimate is consistent with the observations of the HCN and CS lines which require at least such densities for excitation.

The large HCN and CS line widths indicate that the clump is gravitationally unbound. The virial mass of this clump estimated in the usual way (e.g., Zinchenko et al. 1994) is ${M}_{\mathrm{vir}}\sim 30$ M. At the same time, the non-detection of dust emission implies an upper limit for mass orders of magnitudes lower. Therefore, the clump represents a transient entity.

It is worth mentioning the non-detection of HCO+ and N2H+ emission. Both molecules can be destroyed by dissociative recombination (e.g., Zinchenko et al. 2009). Therefore their absence can indicate an enhanced ionization in the dense clump.

5.2. Ionized Gas

In Figure 17 we plot the map of the 15 GHz continuum emission near the SMA1/SMA2 clumps (from the VLA archival data, the angular resolution is about 4 arcsec) which shows the distribution of the ionized gas in this area. This ionized component seems to be associated with the jet traced in the Fe ii emission. The GMRT map at 1280 MHz looks similar but there is a significant positional uncertainty in the GMRT data (Paper I). Properties of the continuum source were estimated in Paper I and by Ojha et al. (2011). The emission measure is EM ∼ (1–2.5) × 107 pc cm−6. Taking into account the observed size of the continuum source, the electron density is n ∼ 3 × 104 cm−3.

6. SURROUNDINGS

The surroundings of the SMA1/SMA2 clumps and high-velocity outflow are traced in several molecular lines. One of the most informative is the N2H+(3–2) transition observed with both the SMA and the IRAM 30 m telescope. The combined map of the N2H+(3–2) emission is presented in Figure 18. It shows an absence of N2H+ in the hot core in accordance with our previous findings (Paper I). The overall morphology of the N2H+ emission suggests that it originates in an envelope around the central cores and the outflow lobes.

The distribution of the SiO(5–4) emission is different (Figure 23). It peaks near the SMA1 core. There is also a feature in the area of the blue outflow lobe that may be associated with the N2H+ emission. The CS(7–6) distribution (right panel in Figure 23) seems to be the most uniform one. It shows a rather smooth, almost spherical halo around the SMA1 and SMA2 cores. There is no sign of the outflow influence on the CS distribution. One may suspect that the optical depth in the CS line is too high and we see only the outer regions of the core. However, the SMA1 and SMA2 clumps are well resolved in this map which makes such explanation less probable. Then, the total gas distribution is apparently not significantly affected, too.

Figure 23.

Figure 23. Left panel: the maps of the SiO(5–4) (left panel) and CS(7–6) (right panel) integrated line emission obtained by combining the SMA and 30 m data (orange contours) overlaid on the continuum image at 0.8 mm. The maps of the CO(3–2) high velocity emission as observed with the IRAM 30 m telescope (blue and red thick contours) from Figure 18 are also shown. The beams for the IRAM 30 m telescope and for the combined maps are shown in the lower left corner of each panel.

Standard image High-resolution image

7. DISCUSSION

The main goal of this study is the characterization of the outflow and probable accretion disk associated with the massive YSO in the S255IR clump. The data presented in the previous sections shed new light on this system.

Concerning the outflow, one of the main findings is that the SMA interferometric data alone give a rather distorted picture. They hint at two highly collimated parallel outflows with different centers of origin. However, this impression is apparently caused by a significant flux loss in the interferometric measurements. The single-dish CO observations clearly show a single less collimated outflow originating at the SMA1 core. The CO emission retrieved by the SMA near the driving source originates apparently from the northern wall of the outflow cavity. This means that this part of the wall is rather thin and bright. The absence of a noticeable CO emission at the opposite side of the wall implies a more diffuse distribution of the emission here, probably due to the density structure of the surrounding medium.

The question arises of how common this effect is in interferometric studies of outflows. We can easily imagine a situation when an interferometric image will show multiple outflows from a single driving source, while in fact there is a single wide-angle outflow.

The observed P–V diagrams for the outflow (Figures 19, 20) most probably indicate periodic ejections from the driving source. Two events can be traced in the data. They apparently created jet knots at different distances from the central star. The older one is responsible for the extended CO outflow. The peaks of the CO emission coincide with bright H2 emission spots (Wang et al. 2011) which probably indicate bow shocks at the heads of the jets. The next ejection event created other bow shocks which are seen in particular in the Fe ii emission (Figure 17). This later ejection entrains dense molecular gas observed in the wings of the HCO+, HCN and CS lines. It is also traced in the CO emission as can be seen in the CO position–velocity diagram (Figures 19, 20). Another manifestation of this activity is apparently the dense high velocity clump (Section 5.1), which most probably represents dense gas at the head of a bow shock. All the jet knots lay practically on a straight line. Therefore the orientation of the jet does not change significantly with time. The age of the first, extended outflow was estimated at about 7000 years by Wang et al. (2011). Assuming the same ejection velocities for the two events and comparing distances of the bow shocks from the driving source we can conclude that the second ejection happened about 1000 years ago. Therefore the time interval between the ejection events is about 6000 years (the age estimates were done under the usual assumption of an inclination angle for a jet of 45°).

Several entrainment mechanisms are usually discussed for molecular outflows (e.g., Arce et al. 2007; Frank et al. 2014). In our case the most probable one is the jet bow-shock model (e.g., Raga & Cabrit 1993). The main argument in favor of this model is the obvious presence of several bow shocks (traced in the H2 and Fe ii emission) clearly associated with the high-velocity molecular gas. This model is also consistent with the observed outflow morphology and kinematics. The shape of the P–V diagram differs from the frequently observed "Hubble law" and shows a range of velocities at the largest distances from the driving source. A similar shape was observed in some other outflows which are apparently driven by bow shocks and this type of shape is expected in the theoretical models (e.g., Lee et al. 2000).

The jet is apparently launched from the accretion disk around the central massive young star. As mentioned above, there is a chain of water masers along the jet. For several water maser spots shown in Figure 5, proper motions have been measured (Goddi et al. 2007). The velocities (from ∼10 to ∼30 km s−1 at a distance of 1.6 kpc) are mostly perpendicular to the jet axis with outward components in some cases. The maser velocity pattern is consistent with clockwise rotation around the core center. Velocities of the masers closest to the center imply a dynamical mass of the order of 20 M. The mass of the central star is estimated to be 27 M (Ojha et al. 2011) assuming a distance of 2.5 kpc. At a distance of 1.6 kpc the mass of the star would be 23.5 ± 3.6 M. This estimate is based on the unpublished value of the star luminosity of (6.8 ± 2.8) × 104 L (for the distance of 1.6 kpc). Now we reconsidered the star SED by including data at millimeter wavelengths from Zinchenko et al. (2009). This results in the luminosity of (3.5 ± 0.3) × 104 L and mass of 20 ± 2 M. It is worth mentioning that Pashchenko et al. (2003) estimated the mass of the central star to be (6–7) M from their analysis of water maser variability. However, they employed a systemic velocity of 8.3 km s−1 that contradicts our findings which clearly show the systemic velocity to be about 4.8 km s−1. In the following we will take the mass of 20 M as the most probable value. This is sufficient to explain velocities of the masers closest to the star. However, maser spots located at larger distances from the center have similar velocities and imply a much larger central mass of the order of 100 M. This value is inconsistent with all other estimates. The corresponding mean density would be unrealistically high. This means that we probably need another explanation for the maser velocities. One speculation is that it might be jet rotation. Rapid rotation of protostellar jets is expected in theoretical models (e.g., Pudritz et al. 2007) and was indeed observed in some cases (Bacciotti et al. 2002).

Water masers are thought to be excited by interaction of the jet with the surrounding medium. The fact that they are moving in one direction on each side of the central star implies some asymmetry in the system. There may be some misalignment between the jet and the rotation axis of the material in the outer parts of the clump (apparently, in the axially symmetric case the observed maser movements would be symmetric relative to the axis). It can be that in the inner part the orientation of the disk plane is different. However, the difference cannot be large. We see also other indications of the jet interaction with the surrounding medium, in particular, increased line widths along the jet.

Figure 5 shows that the water maser condensations are shifted from the jet axis in direction of their proper motion. This projected shift is of the order of 200 AU. At the velocities mentioned above, the maser condensations would travel this distance in about 30–100 years. From this consideration it looks probable that the event that created these masers happened quite recently.

The disk is apparently strongly fragmented as follows from the derived beam filling factor (∼0.2) for CH3CN and CH3OH. The mean gas density in the disk is about 6 × 108 cm−3. The density of the fragments should be much higher. Observations of vibrationally excited HCN indicate densities >1010 cm−3 which is consistent with this picture.

The derived mass of the hot gas (∼0.3 M) is an order of magnitude insufficient to gravitationally bind the core. The virial mass estimated in the usual way (e.g., Zinchenko et al. 1994) from the line width is about 7 M. There can be also some amount of cold material in the disk, as follows from the DCN observations. However, the total mass of the disk is apparently much lower than the mass of the central star which is estimated to be about 20 M (see above). Since the disk mass is significantly lower than the star mass, the disk is probably not self-gravitating and we may expect Keplerian rotation. Unfortunately the achieved angular resolution is not sufficient to reliably measure the disk rotation curve.

For a central mass of 20 M the inclination angle of the disk (derived from a comparison of the observed rotation line-of-sight velocity with the expected velocity for this mass) should be small, about 25° (Figure 6), i.e., the disk is seen almost face-on. This is consistent with our maps which do not show a significant elongation of this object. However, this conclusion looks somewhat suspicious because of the well-collimated appearance of the molecular outflow. Observations at a higher angular resolution are needed to clarify the structure and kinematics of this object.

The temperature of the hot gas derived from the CH3CN, CH3OH, and SO2 observations is about 130–180 K. A similar estimate for the temperature of the hot component was obtained by us many years ago from the analysis of the IRAS data (Zinchenko et al. 1990). The HNCO rotational temperature is somewhat higher, about 320 K. However, as mentioned above, this value probably represents an upper limit to the excitation temperature. In addition, the HNCO excitation can be influenced by the FIR radiation. The HNCO excitation in massive cores was discussed by Zinchenko et al. (2000). That analysis shows that in the present case collisional excitation can be effective. The critical density for excitation of the ${K}_{-1}=3$ ladder is of the order of 1010 cm−3. Such density is quite possible in the SMA1 core as follows from the estimates above. However, radiative excitation by FIR emission via the b-type transitions cannot be excluded. With a column density of $N({{\rm{H}}}_{2})$ ∼ 3× 1024 cm−2 the dust will be optically thick at these frequencies creating a sufficiently strong radiation field. The HNCO excitation especially in the higher ${K}_{-1}$ ladders may reflect the effective temperature of this field.

The highest temperature of the observed molecular material is about 300 K. Taking the source luminosity to be about 3.5 × 104 L (see above) this value corresponds to the dust equilibrium temperature at a distance of 200–300 AU from the star. Since we do not see warmer molecular gas, no dust should be present within this radius in a case of no shielding. However, this radius is comparable to the observed size of the core and we do not see any central hole in the molecular distribution. It is then possible that shielded molecular clumps are distributed at smaller radii.

In this respect an interesting feature of our observations is the apparently relatively cold and rather massive clump with a strong DCN emission. It may be gravitationally bound and the mass of the clump is sufficiently high to consider it as a possible low-mass protostar.

Concerning the surroundings of the SMA1 clump, we see that the outflow strongly affects the chemical composition of the medium. The N2H+ molecules are destroyed along the outflow. The SiO distribution also seems to be significantly affected. At the same time there is no noticeable influence on the CS distribution. The density structure is probably not affected either. As mentioned above (Sections 4.2, 4.4) the nearby clumps are probably influenced by shocks that are apparently associated with this outflow.

8. CONCLUSIONS

We presented the results of our observations of the S255IR area with the SMA at 1.3 mm in the very extended configuration and at 0.8 mm in the compact configuration as well as with the IRAM 30 m at 0.8 mm. The best achieved angular resolution is about 0.4 arcsec. The dust continuum emission and several tens of molecular spectral lines are observed. The majority of the lines is detected only toward the S255IR-SMA1 clump. In summary, our main findings are the following.

  • 1.  
    The S255IR-SMA1 clump apparently represents a rotating structure (probably a disk) around the young massive star. The achieved angular resolution is still insufficient to establish the character (Keplerian or non-Keplerian) of the rotation. The temperature of the molecular gas reaches 130–180 K. The size of the clump is about 500 AU. It is apparently strongly fragmented as follows from the derived small (∼0.2) beam filling factors for various molecules. The mean gas density is about 6 × 108 cm−3. The density of the fragments should be much higher which is confirmed by observations of HNCO and vibrationally excited HCN. The mass of the hot gas is ∼0.3 M and the total mass of the clump is significantly lower than the mass of the central star (about 20 M). The inclination angle of the disk (derived from comparison of the observed rotation line-of-sight velocity with the expected rotation velocity for this mass) should be small, about 25°.
  • 2.  
    We detected a strong DCN(3–2) emission near the center of the SMA1 clump. Most probably it indicates the presence of a rather large amount (≳1 M) of cold (≤80 K) material. This cold clump can be gravitationally bound.
  • 3.  
    High velocity emission is observed in the CO line as well as in lines of high-density tracers HCN, HCO+, CS, and other molecules. The CO outflow is much more extended than that observed in the lines of high-density tracers. Its morphology obtained from a combination of the SMA and IRAM 30 m data is significantly different from that derived from the SMA data alone. The CO emission detected with the SMA traces only one boundary of the outflow and leads to a rather distorted picture of the outflow structure. The velocity of the CO outflow reaches ∼60 km s−1.
  • 4.  
    The outflow is most probably driven by the jet bow shock mechanism. The available data indicate at least two major ejection events with a time interval of several thousand years between them. The direction of the ejections does not change with time. The high-velocity emission in the lines of high-density tracers is associated with the peaks of the Fe ii emission related to the bow shocks caused by rather recent ejections from the SMA1.We detected a dense high-velocity clump apparently associated with one of the bow shocks. It shows a strong emission in the HCN(4–3) and CS(7–6) lines. At the same time there is no detectable HCO+(4–3) or N2H+(3–2) emission which can be probably explained by enhanced ionization.
  • 5.  
    The proper motions of the water masers excited along the jet imply some misalignment of the jet with the rotation axis of the material in the outer parts of the clump. It could be that the orientation of the disk in the inner and outer parts is somewhat different. However, the difference cannot be large.
  • 6.  
    The outflow strongly affects the chemical composition of the surrounding medium. The N2H+ molecules are destroyed along the outflow. The SiO distribution also seems to be significantly affected. At the same time there is no sign of the outflow's influence on the CS distribution. The total gas distribution is apparently not significantly affected either.

This work was supported by the Russian Academy of Sciences (Research program No. 17 of the Department of Physical Sciences), the Russian Foundation for Basic Research (RFBR), the Ministry of Science and Technology (MoST) in Taiwan, and the Department of Science and Technology (DST) of the Government of India in the frameworks of the research grants RFBR 08-02-92001-NSC, RFBR 13-02-92697-Ind, RFBR 15-02-06098, 15-52-45057, and DST-RFBR INT/RUS/RFBR/P-142. S.-Y.L. and Y.-N.S. acknowledge grant support 102-2119-M-001-010-MY2 from MoST. A.M.S. was supported by the Russian Science Foundation (grant No. 15-12-10017). S.V.S. was supported by the Ministry of Education and Science of the Russian Federation (state task No. 3.1781.2014/K). Y.W. was supported by the Swiss National Science Foundation, NSFC 11303097 and 11203081, China. The research is also partly supported by the grant within agreement No. 02.B.49.21.0003 between The Ministry of Education and Science of the Russian Federation and Lobachevsky State University of Nizhni Novgorod and by the Russian Education and Science Ministry Project 3.1252.2014/k. We are grateful to Elena Trofimova for help with line identification and to the anonymous referee for detailed helpful comments. The research has made use of the SIMBAD database, operated by CDS, Strasbourg, France.

Please wait… references are loading.
10.1088/0004-637X/810/1/10