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.

HIGH-RESOLUTION MID-INFRARED SPECTROSCOPY OF NGC 7538 IRS 1: PROBING CHEMISTRY IN A MASSIVE YOUNG STELLAR OBJECT

, , , , and

Published 2009 April 15 © 2009. The American Astronomical Society. All rights reserved.
, , Citation Claudia Knez et al 2009 ApJ 696 471 DOI 10.1088/0004-637X/696/1/471

0004-637X/696/1/471

ABSTRACT

We present high-resolution (R = 75,000–100,000) mid-infrared spectra of the high-mass embedded young star IRS 1 in the NGC 7538 star-forming region. Absorption lines from many rotational states of C2H2, 13C12CH2, CH3, CH4, NH3, HCN, HNCO, and CS are seen. The gas temperature, column density, covering factor, line width, and Doppler shift for each molecule are derived. All molecules were fit with two velocity components between −54 and −63 km s−1. We find high column densities (∼1016 cm−2) for all the observed molecules compared to values previously reported and present new results for CH3 and HNCO. Several physical and chemical models are considered. The favored model involves a nearly edge-on disk around a massive star. Radiation from dust in the inner disk passes through the disk atmosphere, where large molecular column densities can produce the observed absorption line spectrum.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

The NGC 7538 (S158) molecular cloud, located in the Perseus arm at a distance of 2.65 kpc (Moscadelli et al. 2009), contains many protostellar objects, making it a good candidate to study different stages of star formation. Furthermore, there seems to be a progression in evolution from northwest (the visible H ii region, NGC 7538, and IRS 2, an only moderately obscured compact H ii region) to southeast (IRS 9, a deeply embedded core) (Elmegreen & Lada 1977; Campbell & Thompson 1984). The near- and mid-infrared sources IRS 1, 2, and 3 all lie near the northwest end of the star-forming region. Of these, IRS 1 is the brightest mid-infrared and radio continuum source (Martin 1973; Wynn-Williams et al. 1974; Willner 1976). Its exciting star has a luminosity of L > 8 × 104L, which corresponds to a zero-age main sequence (ZAMS) spectral type earlier than O7.5, and an ionizing photon flux of ϕ>1048 s−1 (Werner et al. 1979; Lugo et al. 2004).

IRS 1 is embedded in a dense molecular cloud. Millimeter continuum emission is seen from the surrounding dust (Scoville et al. 1986). Pratap et al. (1989) mapped the HCO+ and HCN emission from within ∼1' of IRS 1 with 3'' resolution. They found evidence for a shell-like structure on scales of 10–20'' around IRS 1. Wilson et al. (1983) observed NH3 absorption toward IRS 1, and Henkel et al. (1984) mapped the absorption with the VLA. The absorption traces warm gas with a high column density of NH3 and is centered near −60 km s−1. The emission, on the other hand, is centered near −56.5 km s−1. The systemic velocity for this source is ∼57 km s−1 (van der Tak et al. 2000). More recently, Zheng et al. (2001) mapped NH3 emission from a more extended region in NGC 7538. This emission probably probes the outer envelope as indicated by the colder temperature and the smaller column density. Using both single-dish and interferometer observations of various molecular tracers, van der Tak et al. (2000) studied the density and temperature structure of both the cold outer envelope and the warm inner material (240–72,000 AU). The inner warm region is characterized by temperatures of a few hundred Kelvin.

Outflows are often present toward protostars, and NGC 7538 IRS 1 is no exception. Campbell (1984) observed IRS 1 with the VLA at 5 and 15 GHz. She found a pair of very compact lobes of continuum radiation, separated in declination by 0farcs2, with emission extending out to ±2''. Her preferred model involves a bipolar ionized outflow from a late O star, collimated by a core of dense gas extending from <65 to >25,000 AU. Further evidence of outflows was found when Gaume et al. (1995) observed IRS 1 with the VLA in the H66α line and the 22 GHz continuum. Their continuum image is similar to that of Campbell, but shows additional structure. They found broad recombination line emission in the two lobes, with a minimum line-to-continuum ratio, which they attribute to high electron density, between the lobes. They propose a model involving a high velocity stellar wind interacting with photoevaporating knots of neutral gas.

More recently, Lugo et al. (2004) have modeled the radio continuum observations of IRS 1 as due to a wind produced by photoevaporation of a circumstellar disk with a radius of 500 AU exposed to UV radiation from the central O star. Maser line emission is seen from H2O (Kameya et al. 1990), OH (Dickel et al. 1982), H2CO (Hoffman et al. 2003), CH3OH (Menten et al. 1986), and NH3 (Madden et al. 1986). Minier et al. (1998) used VLBI observations of the methanol masers to infer the presence of a nearly edge-on rotating disk traced by one of the maser clusters toward IRS 1. The central velocity of this cluster, which is the brightest cluster near IRS 1 in both CH3OH and NH3 masers, is −56.2 km s−1. A second cluster of masers, 0farcs25 to the south, is at −61.0 km s−1. The disk masers lie along a line centered near the gap between the lobes of the free–free continuum radiation, although they are aligned along a position angle ∼30° from the symmetry plane. Pestalozzi et al. (2004) made a detailed model of maser emission from a disk around IRS 1, supporting the suggestion of Minier et al. They concluded that the maser emission comes from disk radii of ∼290–750 AU. However, more recently, De Buizer & Minier (2005) show that the CH3OH masers may trace knots in the outflow instead. Also, De Buizer & Minier identify a disk-like structure perpendicular to the outflow direction in mid-infrared continuum. The extent of the infrared emission is ∼450 AU.

Infrared absorption by dust grains and their icy mantles has also been observed toward IRS 1. Willner (1976) and Willner et al. (1982) observed 9.7 μm silicate absorption and absorption by various ices, including H2O. The ice bands toward IRS 1 are relatively weak, especially in comparison with NGC 7538 IRS 9, which must lie in a colder region of the NGC 7538 cloud. Gibb et al. (2004) observed the complete 2.5–20 μm region with the Short-Wavelength Spectrometer (SWS) on the Infrared Space Observatory (ISO). H2O, CO, and CO2 ices are clearly present, as well as other less securely identified features. Toward NGC 7538 IRS 9, ∼1' southeast of IRS 1, a much richer ice spectrum is seen, with CH3OH, XCN (probably OCN), and CH4 also confidently identified, and OCS, H2CO, HCOOH, and NH3 likely present.

Gas-phase absorption in the infrared has been observed toward IRS 1 by several groups. Mitchell et al. (1990) observed absorption by 12CO and 13CO in their 5 μm v = 1–0 bands. They did not spectrally resolve the lines, and they used a curve of growth analysis assuming pure absorption by Gaussian lines with FWHM of 8 km s−1 to derive the gas temperature and column density from the less-saturated 13CO lines. They found two absorbing components: cold gas at 25 K, and warm gas at 176 K. A gas density of nH > 106 cm−3 was required to maintain the population of the high-J levels in the warm gas. Using the ISO-SWS, Lahuis & van Dishoeck (2000) detected warmer gas than the CO observed by Mitchell et al. They observed the ν5 band of C2H2 at 13.7 μm and the ν2 band of HCN at 14.0 μm. With the SWS resolution of ∼1800, they were not able to resolve individual line shapes, and were most sensitive to the blended lines of the Q branches, but they were able to derive temperatures and column densities from the shapes of the Q branches and the depths of the v = 1–0 fundamental and v = 2–1 hot band Q branches. They derived T = 800 K and N = 0.8 × 1016 cm−2 for C2H2, and T = 600 K and N = 1.0 × 1016 cm−2 for HCN. Boonman et al. (2003) derived T = 500 K for C2H2 from an updated reduction of the spectra. Boonman et al. (2003) and Boonman & van Dishoeck (2003) studied gas absorption from CO2 and H2O using ISO data. They found enhanced abundances toward the inner warm material compared to the cold envelope indicating that grain mantles are sublimating and enriching the gas-phase chemistry close to the protostar. Gas-phase CH4 was also detected toward IRS 9 (Lacy et al. 1991; Boogert et al. 2004) but not IRS 1.

The wealth of observational data available suggest the following physical scenario for NGC 7538 IRS 1. The massive young star is surrounded by a massive cold envelope on scales ∼72,000 AU (van der Tak et al. 2000). At scales below 1000 AU, the temperature increases to ⩾100 K. Inside this radius, there is evidence for a disk (De Buizer & Minier 2005) and small scale knotty outflows (Gaume et al. 1995), both of which can be affected by the UV and X-ray radiation from the protostar (e.g., Stäuber et al. 2004, 2005).

In this paper, we present high-resolution mid-infrared spectroscopy of NGC 7538 IRS 1 showing a rich absorption spectrum containing lines from seven molecules: C2H2, HCN, CH4, CH3, NH3, HNCO, and CS. Mid-infrared absorption is much more sensitive to the inner warm gas than radio observations. Previous infrared studies suffered from limited spectral resolution. Section 2 describes the observations. We then present the model used to derive the column densities and temperatures for the various species in Section 3 and describe the results for each of the molecules in Section 4. Subsequently, in Section 5, we discuss the possible scenario in which we are probing chemistry in a circumstellar disk. In Section 6, we provide some concluding remarks.

2. OBSERVATIONS

2.1. Observations and Data Reduction

The observations were made with TEXES, the Texas Echelon Cross Echelle Spectrograph (Lacy et al. 2002), on the NASA Infrared Telescope Facility (IRTF) in 2001 June and November, 2002 September and December, and 2005 December. TEXES is a high-resolution cross-dispersed spectrograph operating at mid-infrared wavelengths between 5 and 25 μm. It achieves a spectral resolution of R = 75,000–100,000 (Δv = 3–4 km s−1) shortward of 14 μm with a slit width of ∼1farcs5. (Of the observations presented here, only the spectra near 11 μm, which used a 1'' slit, achieved the higher resolution.) The spatial resolution along the North–South oriented slit is ∼1'', slightly larger than the diffraction limit of the IRTF. The spectral and spatial sampling by the 256 × 256 pixel2 Si:As detector array are 1.0 km s−1 and 0farcs35, respectively. At each spectral setting, 5–10 orders of the high-resolution echelon grating are recorded, giving a spectral coverage of Δλtot/λ ∼ 0.5%. At wavelengths shortward of 11  μm full order widths are observed, giving continuous spectral coverage, but longward of 11 μm the order width exceeds the array width, leaving gaps in the observed spectra. The location of the gaps can be shifted in order to optimize line coverage. However, since the gaps are small and, at 13 μm, there are many telluric lines, it would be impractical to repeat observations with small shifts in order to get complete spectral coverage for each setting.

Much of the data were taken in high water vapor conditions, but the observing conditions did not seriously affect the spectra except for increased noise, especially on atmospheric lines. Observations of NGC 7538 IRS 1 were interspersed with observations of a bright asteroid (usually Ceres (greater than 500 Jy), but also Hygeia (∼150 Jy)), which served as a comparison source for removal of telluric absorption features. The telescope was nodded by 3–5'', moving the source along the spectrograph entrance slit to allow subtraction of background emission. At the beginning of each set of nodded observations an ambient temperature blackbody was observed for flat-fielding. The first nods were used to peak up on the source by maximizing the throughput of the slit. The nod pairs taken during peak up were not included in the final sum.

Data reduction followed the standard TEXES procedure (Lacy et al. 2002) of first subtracting the two nod position spectra, flat-fielding with the blackbody-sky difference echellogram (Lacy et al. 1989), and interpolating over spikes and bad pixels. Optimally weighted point-source spectra were then extracted from the echellograms. The spectra were linearized in ln(ν) based on the known optical distortions, and absolute wavenumber calibration was obtained from telluric emission features and corrected for the Earth's motion relative to LSR. The resulting wavenumber scale is correct to within 1 km s−1. The same procedure was used for the asteroid spectra, and then the IRS 1 spectra were divided by the asteroid spectra with correction for differences in airmass. With this procedure, telluric absorption lines as deep as 90% can be removed, although of course the noise increases on lines. However, broad response variations often remained, which were removed by fitting the continuum in each echelon order to a quartic polynomial and dividing. This procedure also sets the continuum in each order to one.

2.2. Line Identification and New Infrared Detections

Ten spectral settings were observed in the 728–820 cm−1 (13.7–12.2 μm) region, two in the 860–930 cm−1 (11.5–10.7 μm) region (see Figure 1), and three in the 1240–1312 cm−1 (8.0–7.6 μm) region (see Figure 2). A total of 110 echelon orders, or ∼70 cm−1, were observed. The original observations were meant to study the line profiles of C2H2 and HCN, which had previously been detected. In addition to the prominent fundamental band lines of C2H2 and HCN, many C2H2 hot band lines were detected in our spectra including lines from the ν4 + ν5 − ν4 and the 2ν5 − ν5 bands. After the hot bands were identified, unidentified lines still remained, some of which were double dipped. The double lines turned out to be CH3 lines which are split due to spin-rotation interactions (see the Appendix). This is the first detection of CH3 toward dense gas and the first ground-based detection of CH3. Previously, CH3 had only been observed with ISO toward the Galactic center (Feuchtgruber et al. 2000). In efforts to detect ethane, C2H6, at 810–820 cm−1, many lines were detected, though none corresponded to C2H6. Some were due to NH3, but many lines remained unidentified (see Figures 3 and 4). Based on their separation we were able to identify the lines as due to HNCO. Interstellar HNCO is a well-known hot core molecule at radio wavelengths (Zinchenko et al. 2000), but it had never been seen at infrared wavelengths. In total, five molecules were observed in the 728–820 cm−1 region: C2H2 (including 13C12CH2), HCN, HNCO, CH3, and NH3. At higher frequencies, CH4, C2H2, and CS were observed in the 1240–1320 cm−1 region. CH4 observations from the ground are possible for this source because of the large Doppler shift with respect to the telluric lines.

Figure 1.

Figure 1. Selected orders (thick black) between 729 and 870 cm−1 are shown with the best fit (red). Positions of lines considered in the fit are given: C2H2 (light blue), 13C12CH2 (blue), hot bands of C2H2 (dark blue), HCN (green), CH3 (orange), HNCO (red), and NH3 (purple). The thin black line indicates the atmospheric transmission. The top panel has a larger scale for the normalized flux (y-axis) to show the lower atmospheric transmission. Note that there are gaps between orders.

Standard image High-resolution image
Figure 2.

Figure 2. Selected orders (thick black) between 1245 and 1285 cm−1 are shown with the best fit (red). Position of lines considered in the fit are given: CH4 (blue) and CS (red). The thin black line indicates the atmospheric transmission. In this region there are no gaps between orders. Instead the orders overlap where the black line is darker.

Standard image High-resolution image
Figure 3.

Figure 3. This figure shows a portion of spectrum containing some NH3 lines as well as a wealth of HNCO lines. These spectral settings led to the identification of HNCO. Figure 4 shows a small portion of this spectrum.

Standard image High-resolution image
Figure 4.

Figure 4. Profiles of several HNCO and NH3 lines in the spectral region shown in Figure 3. The two HNCO subbands shown are R-branch transitions where the K = 1 state is split into upper (U) and lower (L) levels. Because of the close separation the subbands, the lines can overlap for a range of J. See the Appendix for more details on transition rules for HNCO.

Standard image High-resolution image

Table 1 gives a summary of the detected lines. At least five lines were observed for each molecule, with some molecules like C2H2 (including isotopic lines and hot bands), and NH3 having more than 40 lines detected. However, the number of HNCO lines dominates over lines from the other molecules, with over 100 lines detected. Table 2 (full version is available online only) lists all the individual lines detected in our spectra. A selection of spectra including all observed molecules are shown in Figures 1 and 2, along with fits described in Section 3. For most molecules the line strengths were taken from the GEISA03 database.6 The CH3 band strength was obtained from Wormhoudt & McCurdy (1989) and the HNCO oscillator strength was obtained from Lowenthal et al. (2002). The CS oscillator strength was taken from Botschwina & Sebald (1985).

Table 1. Summary of Observed Lines

Molecule Total Number of Detected Lines Band
C2H2 25 ν5
C2H2a 2 ν4 + ν5
C2H2b 20 ν4 + ν5–ν4
13C12CH2c 7 ν5
HCN 10 ν2
NH3 64 ν2
HNCO 125 ν4
CH3 11 ν2
CH4 12 ν24 dyad
CS 6 ν2

Notes. aThese lines were observed at 8 μm. bLines in Q-branch at 731 cm−1. See the top panel of Figure 1. cOnly unblended lines are reported here.

Download table as:  ASCIITypeset image

Table 2. Observed Lines

Molecule Line Wavenumber (cm−1) EJ (K) Band
C2H2 Q(5) 729.289 35.30 ν5
C2H2 Q(6) 729.343 49.42 ν5
C2H2 Q(7) 729.406 65.89 ν5
C2H2 Q(8) 729.477 84.71 ν5
C2H2 Q(9) 729.558 105.89 ν5
C2H2 Q(10) 729.648 129.41 ν5
C2H2 Q(11) 729.747 155.29 ν5
C2H2 Q(14) 730.096 247.02 ν5
C2H2 Q(15) 730.230 282.30 ν5
C2H2 Q(16) 730.373 319.93 ν5
C2H2 Q(18) 730.686 402.22 ν5
C2H2 Q(19) 730.855 446.89 ν5
C2H2 Q(20) 731.034 493.90 ν5
C2H2 Q(26) 732.287 825.20 ν5
C2H2 Q(27) 732.526 888.62 ν5

Notes. Only lines that were individually detected at the 3σ level are included in this list.

Only a portion of this table is shown here to demonstrate its form and content. A machine-readable version of the full table is available.

Download table as:  DataTypeset image

3. SPECTRUM FITTING

To derive physical parameters of the absorbing gas, model spectra were fitted to the observations. For this purpose, the data were divided into two groups. All data at wavenumbers between 700 and 1000 cm−1 (10.7–13.7 μm) were fitted simultaneously, and data at 1240–1280 cm−1 (near 8 μm) were fitted separately. The Marquardt fitting procedure (Bevington & Robinson 2003) was used, which minimizes χ2, the summed squared deviation normalized by the squared noise, between the data and the model.

Fitting was first attempted with a single component for each molecule, but the residuals suggested that a better fit could be obtained with multiple velocity components. The model that we used assumed that each molecule is found in one or two absorbing components. Each component is described by its Doppler shift, VLSR, its Gaussian line width (1/e half width = (2 kT m−1)1/2), b, the molecule's column density, N, the rotational temperature of the gas, T, and the covering factor, C (the fraction of the background continuum source covered by the component). The fitted column density is the average over the partially covered source, so the column density in the covered portion would be N/C. A covering factor of, e.g., 10% means that lines saturate at 90% of the continuum flux and the column density in the absorbing component is ten times the average over the continuum source. The effect modeled in this way could also result from veiling (continuum emission from foreground or surrounding material) or from re-emission in the lines by the absorbing molecules. The components were assumed to overlap by the products of their covering factors (see Figure 5). The following equation was used to calculate the observed transmission:

Equation (1)

where C1 and C2 are the covering factors for the two components. This results in a greater absorption in saturated lines than would result if the optical depths were added first and then the transmission spectrum was calculated from the optical depth spectrum. This approach was chosen because it gave a better fit to the data than adding the optical depths first and assuming the same covering factor, but it assumes that different components absorb along different lines of sight to a partially covered continuum source, which may not be the case.

Figure 5.

Figure 5. Two absorbing components between the continuum source and the observer. Each component covers 10% of the continuum source. The overlap of the two absorbing regions is given by the product of their covering factor. In this case, only 1% of the continuum source is covered by both components.

Standard image High-resolution image

A frequency correction was allowed for each spectral setting observed, to correct for errors in wavelength calibration. With only one exception, the correction was less than a 1 km s−1 Doppler shift. In addition, a broadening of the instrumental resolution was permitted in the fitting to allow for imperfect internal instrument focusing, or possibly a "macroturbulent" broadening in the absorbing gas. The fitting program chose a resolution a factor of 1.2 greater than that derived from gas cell data. A constant frequency resolution (as opposed to a constant Doppler resolution) was used in each of the two fitted regions. This represents well the improving resolving power toward shorter wavelengths in the 10.7–13.7 μm region. In addition to these parameters, the continuum, slope, and curvature of each echelon order were varied, to allow correction of the baseline fitting done during data reduction. The continuum fitting and frequency correction required about 300 free, but rather easily determined, parameters. Fewer parameters are needed to determine the physical conditions such as temperature, column density, line width and covering factor: 70 parameters for the 10.7–13.7 μm spectra and 21 for the 8 μm region. For the 10–13 μm region, we have over 3000 points to constrain the 70 parameters of interest if we characterize the constraining points by the number of lines times the number of pixels for each line.

For C2H2 and HNCO, lines of several bands were observed. For C2H2, in addition to lines of the Q and R branches of the ν5 fundamental, lines of several ν4 + ν5 − ν5 and 2ν5 − ν5 bands (involving absorption from the excited ν4 and ν5 vibration states) were observed. The vibrational temperatures describing the populations of the ν4 and ν5 states were included as free parameters in the fit. R-branch lines of the ν5 band of 13C12CH2 were also observed, and the 13C12CH2/C2H2 abundance ratio was allowed to vary, although it was kept the same in all the C2H2 components. Two C2H2 lines of the ν4 + ν5 band at 7.6 μm were observed as well. For HNCO, lines from the ν4 band were observed. HNCO is an only slightly nonlinear molecule, and its spectrum consists of a series of subbands resembling those of a linear molecule, like C2H2. Its dipole moment oscillates diagonally, giving it a-type subbands, with no change in the angular momentum about its long axis and only P and R branches, and b-type subbands, in which the angular momentum about its long axis changes, and P, Q, and R branches are seen. See the Appendix for further discussion of the HNCO spectrum. Since we did not have laboratory data regarding the relative strengths of the different subbands, we considered them to be caused by different species and summed the abundances to obtain the HNCO abundance. The HCN lines observed were from the ν2 bending mode. We included HCN 2ν2 − ν2 lines in the fit. Although no hot HCN lines were obvious in the spectra, their inclusion lowered χ2 significantly (Δχ2 = 13) and changed the best-fitting NHCN noticeably.

Although it might not be apparent from the spectra, given the shallowness of the lines, many of the detected lines are at least moderately saturated, and some are very saturated. The lines only appear weak because of the small covering factors (or emission filling in the absorption lines) and to a lesser extent the small line widths. This conclusion is very robust; the relative depths of lines of different opacity requires substantial saturation. This is most apparent in the near equality of the depths of the ortho and para lines in the C2H2 Q branch, near 730 cm−1 (see the top panel of Figure 1). HCN in the −60 km s−1 component and NH3 in the −56.5 km s−1 component are so saturated that only lower limits could be placed on their abundances until their isotopomers were included in the fit to provide upper limit constraints. By including isotopomers and intrinsically weak lines, a wide enough range of line optical depths was observed to allow meaningful constraints to be placed on the molecular abundances in spite of the saturation.

The observed lines of CH3 were in the R branch of the ν2 out-of-plane bending mode of this planar molecule. Each of these lines is doubled by the interaction between the electron spin and the molecular rotation (see the Appendix). The CH3 ν2 band center is at 16.5 μm, whereas all observed lines were shortward of 13.5 μm, with J⩾ 7, making us rather insensitive to cold CH3. Lines of the P and Q branches of the NH3 ν2 band were observed. There was no evidence of 15NH3 lines, but they were included in the fit to constrain the NH3 abundance, as was done with H13CN. In the 8 μm region, which was fitted separately, lines of the ν24 dyad of CH4 were observed. Because of the strong absorption by low-J lines of CH4 in the Earth's atmosphere, most of the observed lines originate at J > 4. We were able to observe the R(0) line due to the favorable Doppler shift of the source (∼−57 km s−1). The R(0) line helps constrain the temperature derived from the high J lines. We also detected and fitted absorption by CS.

The derived component parameters are given in Table 3. Uncertainties for the parameters, given in parentheses, are three times the square roots of the diagonal elements of the error matrix, which are uncertainties allowing all other parameters to vary. The noise in the fitted spectra is assumed to be purely photon statistical noise. With this noise estimate, the reduced χ2 ≈ 1.6, indicating that nonstatistical noise sources contribute moderately, or perhaps that a different model for the line profiles is justified. Including nonstatistical noise, the uncertainties given might reasonably be taken to be 95% confidence intervals. The larger fractional errors are asymmetric, and would be more symmetric in the log of the parameters. For column densities, this is a result of the nonlinearity of the curve of growth for moderately saturated lines. Column densities with errors greater than their values are not consistent with zero. For unsaturated lines, the derived covering factors are very uncertain because equivalent widths depend only on the product of the covering factor and the column density. For example, the column densities of HNCO and CH3 have smaller fractional uncertainties than those of C2H2 and HCN, whereas the covering factors have smaller uncertainties for C2H2 and HCN. All of the fitted molecules are detected with very high confidence.

Table 3. Results from the Model Fit to the Spectra

  Component 1 Component 2
Molecule VLSR (km s−1) b (km s−1) Tex (K) N(X) (1016 cm−2) C (×100) VLSR (km s−1) b (km s−1) Tex (K) N(X) (1016 cm−2) C (×100)
C2H2 −55.7 (0.3) 0.6 (0.1) 225 (20) 3.0 (0.6) 5.9 (0.6) −59.4 (0.3) 0.6 (0.1) 191 (10) 2.8 (0.2) 23.7 (1.0)
HCN −56.3 (0.3) 0.8 (0.1) 256 (30) 5.6 (2.8) 8.4 (0.9) −60.0 (0.3) 1.0 (0.4) 456 (127) 1.3 (1.2) 68.4 (49.7)
HNCO −57.2 (0.3) 1.8 (0.2) 319 (27) 0.4 (0.3) 3.8 (0.3) −60.2 (0.2) 0.7 (0.4) 171 (79) 0.1 (0.3) 97.0 (11)
CH3 −54.2 (0.4) 1.8 (0.6) 258 (17) 1.7 (0.2) 97.0 (15) −62.8 (0.5) 1.5 (0.9) 927 (161) 0.6 (0.1) 97.3 (18)
NH3 −57.3 (0.3) 0.9 (1.0) 278 (31) 5.2 (1.5) 8.9 (1.2) −60.1 (0.2) 0.4 (0.1) 248 (32) 2.8 (0.8) 15.8 (3.6)
CH4 −56.3 (0.3) 0.2 (0.04) 674 (360) 20 (8) 5.8 (1.9) −60.1 (0.3) 0.6 (0.3) 668 (250) 16 (13) 99 (13)
CS −55.2 (0.4) 1.0 (0.3) 224 (173) 0.02 (0.03) 100 (8) −59.4 (0.5) 0.7 (0.3) 249 (100) 7.0 (3.0) 5.5 (0.3)

Notes. The column densities listed are Ntot × C.

Download table as:  ASCIITypeset image

4. RESULTS

For most molecules, we found two components with Doppler shifts of VLSR = −56.4 ± 0.3 and −59.7 ± 0.3 km s−1, with widths of b = 1–2 km s−1 (see Table 3). Most of the component temperatures were in the range 200–400 K and covering factors were typically 0.04–0.08, with C2H2 requiring somewhat larger values of 0.06 and 0.18. HNCO has somewhat broader component line widths but other parameters are similar. CH3 has the broadest component line widths, with a greater velocity spacing between the two components and the hottest component. Both CS and CH3 are fitted with much larger covering factors (consistent with 100%) than were the other molecules. However, these covering factors are quite uncertain due to the weakness of the observed lines, and the large values may not be significant. The model spectra resulting from the best fit parameters can be seen in Figures 1 and 2.

In addition to agreeing among the different molecules (except CH3), the Doppler shifts agree with the two most commonly observed values for molecules seen at radio wavelengths. This gives us some confidence that the two velocity components, at −56.5 and −60 km s−1, are real, although whether they are two separate components, or just a way of describing a non-Gaussian velocity distribution is difficult to determine from the data. Generally, the temperatures of the two components for a given molecule are not very different, although a fit constraining them to be equal was significantly worse than the fit allowing all temperatures to vary. Line widths, column densities, and covering factors vary more. The fit was very poor when they were constrained to be equal for all molecules.

Note that our quoted column densities are averages over the lines of sight to the continuum emitting material, or column densities in the absorbing gas multiplied by the covering factors. There are several possible interpretations for the covering factor parameters: the continuum source could in fact be partially covered, foreground or surrounding emission could veil the spectrum, or re-emission in the lines could fill in the absorption. In all of these cases, the column densities through the absorbing material are given by N/C, rather than N, the average column density over the line of sight in the partial covering interpretation. Consequently, if we could determine the H2 column density along the lines of sight on which we see absorption we should compare that to our N/C values.

4.1. H2 and CO From Previous Measurements

Before discussing the results of our measurements of individual molecules, we attempt to derive H2 and CO column densities from previous measurements. The H2 column toward NGC 7538 IRS 1 has not been directly measured, but it may be estimated from the extinction measurements of Willner (1976) with the assumption of a normal interstellar gas-to-dust ratio. This method gives ${\rm N}_{\rm H_2} = 6\hbox{--}9 \times 10^{22}\,{\rm cm}^{-2}$. If the continuum source is nonuniformly covered, this column density represents an average over the source. Hence, it should probably be compared to the values we derive for average column densities, N, of the molecules we observe. However, the silicate optical depth includes extinction from all material along the line of sight to NGC 7538, most of which is unlikely to contain our molecules. This suggests that 6–9 × 1022 cm−2 is an overestimate of ${\rm N}_{\rm H_2}$ in the relevant gas. On the other hand, we suggest below that we may be observing absorption by gas in the atmosphere of a protoplanetary disk, and dust in such a region could be depleted by settling toward the disk midplane. This effect would result in an underestimate of ${\rm N}_{\rm H_2}$ from the extinction. We give the ratios of our derived column densities to an H2 column density of 7.5 × 1022 cm−2 in Table 4.

Table 4. Abundances with Respect to CO and H2

Molecule N(X)/N(CO)a (10−3) N(X)/N(H2a (10−7)
C2H2 5.7 7.7
HCN 6.9 9.2
HNCO 0.5 0.6
CH3 2.3 3.1
NH3 8.1 10.8
CH4 36 48
CS 7.0 9.3

Notes. aN(CO) = 1.0 × 1019 cm−2 and N(H2) = 7.5 × 1022 cm−2. See the text for discussion of whether these molecules are measured on the same lines of sight as the ones we observe. We use N listed in Table 3.

Download table as:  ASCIITypeset image

It may be preferable to compare our column densities to the column density of CO, which has been observed in absorption toward NGC 7538 IRS 1 by Mitchell et al. (1990). They obtain a column density of 1.5 ± 0.3 × 1017 cm−2 for 13CO in cold gas at 25 K and 1.4 ± 0.1 × 1017 cm−2 in warm gas at 176 K. We would not be sensitive to their cold gas. Presumably their warm gas includes both of our components, although our temperatures are generally higher than theirs. If we assume 12CO/13CO = 72, using the 12C/13C ratio we derived from our C2H2 observations, we derive the abundances of our molecules relative to CO listed in Table 4. (Note that Mitchell et al. use 12C/13C = 89.) However, we should note that we have measured one spectral setting including CO lines, while searching for OCS absorption near 5 μm. The noise is rather high, but it is apparent that the 13CO lines are not well described by moderately saturated Gaussians, as assumed by Mitchell et al. The 13CO lines are deeper and broader than the lines of any of our other molecules, and have prominent blue-shifted shoulders probably tracing an outflow. We suspect that the lines are more saturated, and the CO column density larger than Mitchell et al. concluded. The different shapes and greater depths of the CO lines than our C2H2 lines also indicates that CO probes different gas. Likely, C2H2 and the other molecules we observe in the mid-infrared are found only in unusual regions, whereas CO is distributed along the entire line of sight.

In general, we conclude that the ratios of column densities to those of either H2 or CO given in Table 4 are quite uncertain because the different column density tracers may be sensitive to different components along the line of sight. The column densities of the different molecules we observe should be much more comparable.

4.2. C2H2 and HCN

Since our initial search for molecules began with C2H2 and HCN, the first models only included those two molecules. In these models, we derived similar temperatures and column densities as derived from ISO observations (Lahuis & van Dishoeck 2000; Boonman et al. 2003). Once we include absorption from other molecules, HNCO in particular, the temperatures for HCN and especially for C2H2 are lowered significantly. Some HNCO lines overlap high-J lines of C2H2, which means that more hot C2H2 is necessary to produce the observed absorption when HNCO is not included in the model. In our final model including all of the observed molecules, our derived column densities of C2H2, N = 5.8 × 1016 cm−2, and HCN, N = 6.9 × 1016 cm−2, (see Table 3) are larger than those derived in Lahuis & van Dishoeck, $N_{\rm C_2H_2} = 8 \times 10^{15}$ cm−2 and NHCN = 1.0 × 1016 cm−2. Part of the difference is due to the different Doppler parameter used: b = 5 km s−1 is used in Lahuis & van Dishoeck, whereas, we derive b ∼ 1 km s−1. We derive lower temperatures, $T_{\rm C_2H_2} = 190\hbox{--}230$ K and THCN = 250–450 K than derived by Lahuis & van Dishoeck, $T_{\rm C_2H_2} = 800$ K and THCN = 600 K, and Boonman et al. (2003), $T_{\rm C_2H_2} = 500$ K. We also find that we need a small covering factor to account for the saturated yet shallow lines. The level of saturation is difficult to determine from the ISO data. In order to verify that our fit to the TEXES data is also consistent with the ISO observations, we take our model at the ISO-SWS resolution and compare to the data (Figure 6). The model matches the fundamental Q-branches for the two molecules as well as a hot band Q-branch of C2H2.

Figure 6.

Figure 6. Spectrum of NGC 7538 IRS 1 as seen by ISO is shown in black. A fit to the C2H2 and HCN fundamental Q-branches as well as a hot band Q-branch of C2H2 is shown in grey. The parameters are set from the TEXES high-resolution observations. The models gives a good fit to the ISO spectrum.

Standard image High-resolution image

C2H2 is expected to be in rotational local thermal equilibrium (LTE) since it has no permanent dipole moment. The fact that the two excited vibrational states of C2H2 from which absorption was observed, ν4 and ν5, are found to be populated near LTE at the C2H2 rotational temperature, or even slightly higher, provides information about the physical conditions of the gas. For both bands, the hot bands indicate they are slightly out of LTE since the vibrational temperatures (Tv) are higher than the rotational temperatures (Tr). For the ν4 + ν5 − ν4 bands, the Tv temperatures for the two components were 346 and 271 K while for the 2ν5 − ν5 bands, they were 277 and 227 K. The rotational temperatures for the ν5 fundamental are 225 and 191 K for the two components. In LTE, the vibrational temperatures for the hot bands would equal the rotational temperature measured by the fundamental band lines.

The ν4 level has no allowed radiative transitions to the ground vibrational level, and none of the rotational levels of the ground vibrational level are coupled radiatively, in both cases due to the symmetry of the C2H2 molecule. This suggests that the populations of the ground and ν4 states should be set by collisions, and so should be in LTE at the same temperature. However, the ν4 level can be radiatively pumped by absorption in the 7.6 μm ν4 + ν5 band followed by emission of photons in the ν4 + ν5 − ν4 band. The ν5 state must be populated predominantly radiatively for any plausible density, since the critical density is >1010 cm−3. From these considerations, we conclude that the brightness temperature of the radiation field seen by the observed gas must be close to the kinetic temperature of the gas. In addition, the fact that the measured rotational temperature for HCN is similar to that of C2H2 indicates that the rotational levels of HCN are kept populated either by collisions, which would require a gas density >107 cm−3, or by infrared pumping through the ν2 transitions. Lahuis & van Dishoeck (2000) conclude that either excitation mode is possible. The presence of C2H25 − ν5 absorption requires infrared pumping, although ν4 + ν5 − ν4 absorption does not.

We can attempt to estimate the extent to which emission in our lines fills in the absorption, at least for C2H2, by using the 2ν5 − ν5 lines as probes of the ν5 population. Our spectra require a ν5 vibrational temperature ∼300 K. We need to compare this number to the blackbody brightness temperature of the background continuum radiation. From the shape of the 8–13 μm spectrum, Willner derived a dust temperature of 330 K assuming the emitting dust is optically thick, or 370 K assuming optically thin silicate emission. He preferred the latter model, which gives a brightness temperature only when combined with an assumption about the solid angle subtended by the source. In either case, our vibrational temperature is only ∼10% less than the background temperature, so that the source function is ∼30% below the background intensity, and lines would be expected to saturate at ∼70% of the continuum. However, the continuum spectrum from which Willner derived a dust temperature is the average over the continuum source. If the gas we observe covers only a small fraction of the source, the brightness temperature of the radiation passing through that gas could be much larger than the average.

Given the many assumptions and uncertainties in this calculation, it is difficult to make a strong statement. The level at which our C2H2 lines saturate, ∼85% of the continuum, which we model with a small covering factor, could instead be due partly or entirely to emission by the absorbing gas. It is possible for the lines to saturate at 85% of the continuum even with small covering factors (e.g., 24% and 6% for C2H2) since the line widths are narrower than our spectral resolution. Presumably the same statement applies to the other observed molecules.

4.3. NH3

From our infrared observations, we find that NH3 has two components: one at −57.3 km s−1 and the other at −60.1 km s−1. Most of the NH3 is found in the first component, for which the temperature is ∼278 K and the column density is 5.2 × 1016 cm−2. The second component has a lower temperature (∼248 K) and the column density is 2.8 × 1016 cm−2(see Table 3). NH3 absorption toward IRS 1 has also been studied in the radio by Wilson et al. (1983) and Henkel et al. (1984). They found $N_{\rm NH_3} = 2 \times 10^{18}$ cm−2 in gas with T = 170–220 K and VLSR = −60 km s−1. Although their temperature and Doppler shift is in reasonable agreement with our bluer absorption component, their column density is nearly 20 times ours. Without constraining the 15N/14N isotopic ratio, our data allow an NH3 column as large as that from radio observations, but with this NH3 column 15NH3 lines should have been apparent in our spectra. It is possible that we sample only a fraction of the column observed at cm wavelengths as a result of larger dust opacity at mid-infrared wavelengths preventing us from observing lines of sight through the densest regions of a knotty gas distribution. It may be notable that Wilson et al. found, as did we, that although relative depths of lines (in their case hyperfine components) require large optical depths, line shapes do not appear flat-bottomed as would be expected if they are optically thick. They suggest that the absorbing gas may be in small spectrally and spatially unresolved knots, each of which is optically thick, and which combine to make a line that does not have the shape of a thick line. Our use of two components in our fit may be a way of approximating this situation.

4.4. HNCO

The velocities for the two components are −57.2 and −60.2 km s−1, with corresponding temperatures of 319 and 171 K and column densities of 4 × 1015 and 1 × 1015 cm−2. Since the HNCO temperatures and velocities derived from our observations agree with the values for the other molecules, we can assume that HNCO is in the same gas. It appears the region probed by HNCO is very small in angular size. Zinchenko et al. (2000) detect HNCO emission in the radio and derive a beam-averaged column density of 1.1 × 1014 cm−2. This value is an order of magnitude lower than our observed column density. However, beam dilution in the radio observations can explain the difference in the column densities.

4.5. CH3

CH3 is detected for the first time toward warm, dense gas. The parameters derived for this molecule do not agree as well with the other molecules, although the weakness of the observed lines makes the derived parameters rather uncertain. The separation between the centroid velocities for the two components, VLSR = −54.2 and −62.8, is nearly twice the separation found for the other molecules. Also, the covering factor for both of the components is consistent with 100%, whereas as it is closer to 10% for the other molecules. In addition, the temperature of the second component is much hotter (greater than 900 K) than the temperatures derived for the other molecules. We note though that the fact that the observed lines are of high excitation makes them particularly sensitive to a hot gas component, which may not have been noticed for other molecules. And the weakness of the lines makes the derived covering factor (which is inferred from relative line depths indicating saturation) very uncertain.

4.6. CH4 and CS

CH4 gas-phase absorption had been observed toward the neighboring protostar IRS 9 (Lacy et al. 1991; Boogert et al. 2004) but had not previously been seen toward IRS 1. We find centroid velocities of −56.3 and −60.1 km s−1, which are similar to the other molecules. The temperatures are higher than for most molecules (∼670 K for both components) but have large uncertainties (see Table 3). The best fit seems to indicate that one of the components covers the entire source whereas the other covers about 6% of the source. Of all of the observed molecules, CH4 has the highest column density. The implications of the high CH4 column density the chemistry are discussed in Section 5.1.

We detect only six lines of CS (see Table 2—full table available online only). The lines are rather weak and parameters are poorly constrained. Despite uncertainties, the parameters agree with those for other molecules. The centroid velocities are at −55.2 and −59.4 km s−1. The temperatures, T1 = 224 K and T2 = 249 K, are similar to those found other molecules (except for CH4 and CH3). From radio studies of the envelope around IRS 1, a CS abundance of ∼10−10 is found (Plume et al. 1997; Mueller et al. 2002; Shirley et al. 2003). However, also using radio observations, van der Tak et al. (2000) derive a CS abundance of ∼10−8. Our abundance is about 2 orders of magnitude larger than the largest estimate of the abundance in the envelope indicating that our observations are probing material closer to the protostar.

5. MODELS

5.1. Chemical Models

We now consider the implications of the observations for the chemistry of the observed material. Hot core models can be used to represent two of the scenarios discussed in Section 1: (1) material in a circumstellar disk and (2) photoevaporating knots of neutral molecular material. Hot cores are small (r < 0.1 pc), dense ($n_{\rm H_2} > 10^7$ cm−3) and hot (T > 100 K) regions associated with high-mass young stellar objects (Kurtz et al. 2000). Chemically, hot cores are identified by an enhanced abundance of fully hydrogenated molecules such as H2O and NH3, which are usually observed through rotational transitions at submillimeter and millimeter wavelengths. These abundances are enhanced with respect to the cold molecular envelope. A hot core is thought to form as a result of the protostar heating nearby material, which triggers ice sublimation from the grain mantles (Charnley et al. 1992). HNCO is another hot core molecule which has been found in the Orion hot core with an abundance ∼10−8 relative to H2 (Zinchenko et al. 2000).

Gas-phase HNCO is possibly coming from evaporation of grain mantles. The 4.62 μm feature observed toward some protostars has been attributed to OCN frozen in grain mantles (e.g., Pendleton et al. 1999; van Broekhuizen et al. 2004). Recent ISO results of solid state features toward protostars show an OCN upper limit of 1016 cm−2 toward IRS 1 (Gibb et al. 2004). The total HNCO column derived in this work is 5.4 × 1015 cm−2, a factor of 2 lower than the solid OCN limit. In general, IRS 1 does not show many ice features compared to the colder nearby source IRS 9 (see Gibb et al. 2004), which has a column density N(OCN) = 1.2 × 1017 cm−2. If solid OCN sublimates and captures a proton to form HNCO, the expected column density for HNCO would be comparable to the observed OCN ice toward cold lines of sight (like IRS 9). However, the gas-phase column observed is at least an order of magnitude smaller than that of the solid state ion toward sources like IRS 9. This indicates that when OCN sublimates, a fraction of the ices goes into forming HNCO, while most of HNCO is destroyed in the warm gas-phase chemistry on short time scales. Further studies of how OCN reacts in the gas-phase are needed in order to understand the difference between ice and gas phase column densities.

Similarly, CH4 is probably also evaporating from the grain mantles, and thus the gas-phase abundance is highly enhanced compared with the cold molecular envelope. Studies of the solid CH4 content toward IRS 1 show that the column density is 14 times less than our observed gas-phase column (Gibb et al. 2004), suggesting that the solid CH4 on grains has sublimated. In comparison, IRS 9 has a higher content of solid CH4 than gas-phase CH4 (Boogert et al. 2004; Gibb et al. 2004). If we consider that IRS 1 is likely a more evolved object than IRS 9 (Elmegreen & Lada 1977), the higher gas-phase CH4 toward IRS 1 also points to grain mantle evaporation. The high abundance of CH4 means that daughter products such as CH3 and CH2 should be abundant. While the photodissociation of CH4 preferentially forms CH2 rather than CH3, our observations of CH3 agree with this branching ratio. We predict that CH2 is also present toward IRS 1 with a higher abundance than CH3. It is uncertain where C2H2 is formed in the gas-phase chemistry or is evaporating from grain mantles. However, most models do not predict the observed column densities from formation in gas-phase chemistry only.

From our observations we see a high abundance of the parent molecules NH3 and CH4. High abundances of the other molecules are also observed. Nomura & Millar (2004) present a time dependent chemical evolution of molecules in a hot core. The models begin at the time when the protostar turns on (t = 0 years). Table 5 shows the abundances with respect to CO at an age of 104 years. The hot core model seems to underpredict abundances with respect to CO of all our observed molecules. However, as mentioned in Section 4.2, the CO column density found by Mitchell et al. (1990) may not be the appropriate number with which to compare our observations. If the CO column along the lines of sight containing our molecules is larger than the column measured by Mitchell et al., then perhaps the abundances from the model will more closely resemble the observations. We can also compare the relative abundances between our observed molecules, avoiding the uncertainties in whether the CO and dust absorption trace the lines of sight that our molecules are found in. The observed and predicted values for N(CH3)/N(CH4) are similar, ∼0.01. For the other molecules the ratios from the models at an age of 104 years are very different from the observed values. The values for the models assume that we can see all the material and comparing the model results directly with the observations without accounting for line-of-sight does not make for a good comparison. Figure 7 shows the column density evolution for six of the detected molecules as determined by Nomura & Millar (2004) corrected for the observed beam (H. Nomura 2008, private communication). For each molecule, the observed range of column densities is shown in the horizontal shaded area. This hot core model can predict the observed column densities for four of the six molecules plotted. However, the ages indicated by each molecule do not give a consistent overall age. For CH4 and HCN the model overpredicts the observed column densities. The other molecules suggest ages between 2 × 103 and 2 × 106 years. While ages vary by three orders of magnitude, it should be noted that the model has not been optimized to be an accurate representation of this source. With chemical models that more closely represent the source, it may be possible to determine a more consistent age.

Figure 7.

Figure 7. These plots illustrate the determination of ages based on the column densities of the various molecules compared with time dependent chemical model. The model data are from Nomura & Millar (2004). The horizontal lines indicate the range in column densities derived in this paper. The vertical lines indicate the age range corresponding to those column densities. Note that for CH4 and HCN the model overpredicts the column density at all ages.

Standard image High-resolution image

Table 5. N(X)/N(CO) from Chemical Models

Molecule Observed (10−3) Hot Core Modela (10−3) Disk Modelsb (10−3)
C2H2 5.7 0.88 0.06–0.1
HCN 6.9 0.02 16–24
CH3 2.3 0.016 0.02–1.6
NH3 8.1 0.76 1–6
CH4 36 1.6 7–35
CS* 7.0 0.0031 0.03–0.07

Notes. The model values were taken at 104 years. aNomura & Millar (2004) and H. Nomura (2008, private communication). bNguyen et al. (2002).

Download table as:  ASCIITypeset image

If we consider the disk scenario, we can also compare our observations to a chemical model of a photoevaporating disk. Nguyen et al. (2002) present a chemical model for a disk around a 10 M star. The model consists of a flared disk with a photoevaporating layer on the surface. Table 5 compares our derived abundances with respect to CO for the various molecules to the predicted values from disk models. The summarized results of models taking the midplane temperature for the temperature of the layers below the photodissociation region (PDR) are discussed here. The difference among the three models is the variation in the X-ray ionization rate, ζ (1ζ = 1.3 × 10−17s−1), from 1 to 105ζ. Table 5 shows that the abundance for C2H2 is underpredicted. In the chemical model, C2H2 is produced only by gas phase reactions. One possible explanation is that C2H2 is formed as ice on grain mantles. The abundance of NH3 is also underpredicted. The model with the highest X-ray ionization rate produces CH3 abundances close to the observed value. For that model, N(CH3)/N(CH4) is 4 times larger than the observed ratio. These models do not fit our observations very well but can serve as a guide to the expected abundances in such a scenario. The model by Nguyen et al. has not been modified to try to fit data. So, adapting the model for this object may give predictions closer to the observations. The star in NGC 7538 IRS 1 is believed to be a 30 M star. The different radiation field may affect the resulting chemistry.

Doty et al. (2002) present a model of chemical evolution of the envelopes of massive protostars using AFGL 2591 as an example. From that model, they conclude that enhancements of C2H2, HCN, and CH4 are possible at late times in the evolution (greater than 105 years) and at high temperature (T∼ 800 K). Our derived temperatures indicate cooler material, yet the molecules seem to be enhanced in the line of sight. Alternatively, these species could be produced at high temperature in the inner disk, and subsequently brought outward by the disk wind. Including UV and X-ray radiation in the model affects the chemistry especially relating to HCN. Stäuber et al. (2004, 2005) find that increasing the UV and X-ray flux can lead to enhanced HCN starting by the ionization of N2. The column densities for HCN and CS from the Stäuber et al. model of protostar AFGL 2591 agree with our observed abundances while their column densities for C2H2, CH4, and NH3 are lower than our observed values.

Hot core and disk chemistry models predict the enhancement of molecules such as C2H2, CH4, and NH3. However, the observed abundance of C2H2 is higher than what models predict from warm gas-phase chemistry (see Table 5). This indicates that C2H2 is probably frozen on dust grains and sublimates along with molecules like CH4. Boudin et al. (1998) studied the solid features of C2H2, especially when mixed with H2O or CO. They found that the C2H2 features broaden substantially when mixed with H2O and to a lesser extent when mixed with CO. They compare the laboratory data to ISO data for the colder neighbor, IRS 9, which resulted in an upper limit of 8 × 1017 cm−2 for the column density of solid C2H2. This upper limit is consistent with the observed gas-phase column density seen toward IRS 1. So, it is possible that solid C2H2 is sublimating from grain mantles as protostars heat the environment. The observed CH3/CH4 ratio agrees with the branching ratio for the destruction of CH4. Based on these results CH2 should also be very abundant.

5.2. Physical Models

We now attempt to construct a physical and geometrical model of NGC 7538 IRS 1. Within the possible scenarios presented from various radio and infrared observations (e.g., Minier et al. 2001; Lugo et al. 2004; De Buizer & Minier 2005; Kraus et al. 2006), we propose a scenario in which the molecular absorption presented here comes from a circumstellar disk. We will examine other possibilities first. Figure 8 depicts the possible scenarios allowed by the available observations.

Figure 8.

Figure 8. A cartoon representation of the IRS 1 region. The absorbing material containing the molecules observed can either be located in the disk or in the hot knots, while the continuum is coming from the photoevaporating layers of the the inner rim and surface of the disk. Flaring is not to scale. The continuum can also be from other parts of the disk for which lines of sight do not cross the material with our molecules. There may also be more extended dust emission from the envelope that contributes to the continuum but does not contain our molecules in high abundances. In this figure, North roughly points toward the top of the page.

Standard image High-resolution image

We can quickly rule out a simple model in which the absorbing molecular gas is in the foreground molecular cloud and not closely associated with IRS 1. Our observations require the gas to be much hotter (T∼ 300 K) and denser ($n_{\rm H_2} \sim 10^7$ cm−3 to maintain rotational LTE of HCN out to J = 21) than is found away from luminous sources in molecular clouds. According to van der Tak et al. (2000), temperatures do not reach the observed values until you get to within 400 AU of the central star. In addition, the infrared radiation field must have a brightness temperature ∼300 K to populate the C2H2 ν5 level sufficiently to account for the observed 2ν5–ν5 absorption. Undoubtedly, the absorbing molecular gas is in close proximity to the infrared continuum source IRS 1.

It is not quite so easy to rule out a model in which the absorbing molecules are in boundary region between the envelope around the IRS 1 hypercompact H ii region and the outflow. This gas could be compressed by the ionized wind, and if it is as close as 0farcs1, or 280 AU, from a 105L source it would have a temperature near 300 K. The temperature structure determined by van der Tak et al. (2000) indicates that indeed temperatures range from 200 to 400 K at radii of 220–380 AU. However, interaction with the ionized wind, which has a velocity ∼100 km s−1, would be expected to accelerate the gas, causing broad, blue-shifted absorption (see van der Tak et al. 2000). In contrast, the observed lines have widths of <8 km s−1 and have centroids within a few km s−1 of those seen in surrounding molecular gas. Another argument against the presence of the observed gas in an envelope around IRS 1 is the fact that net absorption is seen. Since the vibrational temperature is comparable to the brightness temperature of the continuum radiation, emission lines would be seen if the molecular gas had a larger extent than the continuum source. If the lines arose in molecular shell surrounding the hypercompact H ii region, this probably would be the case.

On the other hand, Campbell (1984) proposes that the centimeter continuum from IRS 1 results from partially ionized material in an outflow. The centimeter continuum emission has been spatially resolved into knots by Gaume et al. (1995), who suggest that the emission comes from photoevaporation of knots of neutral molecular material (see Figure 8). The CH3OH masers seen by Minier et al. (2000) also seem to trace the knotty structure (in addition to the disk described in Section 1) probed by the cm continuum emission. The knots may be material stripped from the disk. The stellar wind would then blow on these knots. However, if our molecules are in these knots we would expect large velocity difference between the knots resulting in broad lines, which are not observed. If we compare the velocities we derive for the two components of our molecules to those observed for the CH3OH masers, we find that they match the two velocities found for the A cluster of masers. This indicates that the infrared observations may be probing the same gas as the CH3OH masers. This scenario is further supported by Kraus et al. (2006). However, Kraus et al. also find that the interpretation of the masers tracing a disk also agrees with their observations.

We propose that the absorbing material is in a circumstellar disk. There is evidence for a disk close to edge-on from other observations and models (e.g., Pestalozzi et al. 2004; Lugo et al. 2004; De Buizer & Minier 2005). In this situation, hot dust in the surface layers or the inner rim of a flared disk could provide the continuum source. With a nearly edge-on orientation, radiation from the inner region could pass through the outer disk atmosphere, where absorption lines could be formed. Emission from some of the dust on the surface of a slightly inclined disk or in an outflow could be observed directly, accounting for the dilution of the spectrum that we modeled with a small covering factor. This scenario is similar to the one described by Lahuis et al. (2006) to describe C2H2 and HCN absorption toward the low-mass protostar IRS 46 in Ophiuchus (YLW 16B). Detailed modeling of the disk structure of IRS 46 showed that it was possible to have the absorbing material in the disk illuminated by the continuum from the surface. Figure 8 is a cartoon of our proposed picture, where the orange rays represent the continuum. Some of the continuum radiation passes through the absorbing outer disk. In this scenario, dust settling is likely to occur, allowing for observations of large columns of gas without too much dust.

6. CONCLUSIONS

Here we present a summary of our conclusions.

  • 1.  
    NGC 7538IRS 1 has a rich spectrum in the mid-infrared. At high spectral resolution, lines of previously observed molecules such as C2H2 and HCN are seen as well as weak lines of molecules such as HNCO, CH3, NH3, CH4, and CS. We have presented the first infrared detection of interstellar HNCO. This is also the first detection of CH3 toward dense gas.
  • 2.  
    The data show shallow yet saturated lines indicating that either the absorbing gas does not fully cover the continuum source or there is emission filling in the absorption. The observed ortho:para equivalent width ratio for C2H2 is ∼1.5 instead of the expected value of 3.
  • 3.  
    The rotational temperatures for the various molecules range between 200 and 400 K with the exception of CH3, which has one hot component (T ∼ 900 K). From the derived temperatures, we find that the gas is within ∼400 AU of the central star (van der Tak et al. 2000).
  • 4.  
    We suggest that the material traces a close to edge-on circumstellar disk as determined from other observations (e.g., Minier et al. 1998). However, it can also be tracing hot, knotty material close in to the star (see Figure 8).
  • 5.  
    Because it is possible to use hot core chemistry to describe chemistry in a disk, the different chemical models do not help identify the physical location of the absorbing gas. However, the models do seem to indicate that NGC 7538 IRS 1 is an evolved protostar with an age ∼105 years.
  • 6.  
    Our observations help constrain the chemistry in massive protostars. The abundances for all the molecules are enhanced, except for possibly CS (see discussion in Section 4.6). Species known to be present in icy mantles such as CH4 and NH3 have high column densities suggesting that they have recently sublimated from grain mantles. The high abundance of the daughter molecules such CH3 gives constraints on the gas-phase chemistry happening after sublimation of ices. HNCO also provides some constraints about the presence of OCN on dust grains. It is unclear whether C2H2 is a parent molecule or a product of warm gas-phase chemistry. For molecules like HCN, UV/X-ray radiation is important in producing the observed abundances.

C.K. thanks A.C.A. Boogert, A.M.S. Boonman-Gloudemans, S. Doty, and F. Lahuis for useful discussions. The authors thank T.K. Greathouse for assisting with the observations. C.K. acknowledges support from the NASA Astrobiology Program under RTOP-344-53-51. This work was partly supported by NSF grant AST-0607312.

APPENDIX: SPECTROSCOPY OF CH3 AND HNCO

CH3 is a nearly planar molecule and thus we do not expect to see inversion transitions as observed in NH3. We observe the out-of-plane vibrational mode (similar to the umbrella mode in NH3). We observe splitting due to spin-rotation interactions. For symmetric-top molecules, the splitting of a rotational level is given by

Equation (A1)

where $\epsilon ^{^{(\nu)}}\!\!\!\!\!_{_{bb}}$ and $\epsilon ^{^{ (\nu)}}\!\!\!\!\!_{_{cc}}$ are the spin-rotation coupling constants for the B and C axes, respectively, at a given vibrational state, ν. We use Equation (A1) in deriving the splitting for the lines we observed. The values for $\epsilon ^{^{(\nu)}}\!\!\!\!\!_{_{bb}}$ and $\epsilon ^{ ^{(\nu)}}\!\!\!\!\!_{_{cc}}$ were taken from Yamada et al. (1981).

HNCO is a quasi-linear, nearly symmetric-top molecule. The moment of inertia about the figure axis is small and the moments of inertia about the other two axes are much larger (∼100 times larger) and about equal to each other. Because it is a slightly asymmetric molecule, the levels with K ≠ 0 are split into two components. Steiner et al. (1979) refer to these levels as upper (U) and lower (L). Table 6 shows the selection rules for HNCO.

Table 6. HNCO Selection Rules

ΔK = 0   ΔK = ±1
K ≠ 0        
ΔJ = 0 ΔJ = ±1   ΔJ = 0 ΔJ = ±1
UL UU   UU UL
  LL   LL  

Download table as:  ASCIITypeset image

In Figure 4, the HNCO lines labeled R(J)1U correspond to lines where ΔK = 0 and ΔJ = 1 and K = 1 upper. Likewise, the lines labeled R(J)1L are lines where ΔK = 0 and ΔJ = 1 and K = 1 lower. Because the splitting between the two levels is small the lines overlap in the spectral region shown in Figures 3 and 4. Some lines from the upper and lower K states are blended together and give the appearance of double peaked lines.

Footnotes

Please wait… references are loading.
10.1088/0004-637X/696/1/471