Brought to you by:

A publishing partnership

THE INITIAL CONDITIONS OF CLUSTERED STAR FORMATION. I. NH3 OBSERVATIONS OF DENSE CORES IN OPHIUCHUS

, , , and

Published 2009 May 13 © 2009. The American Astronomical Society. All rights reserved.
, , Citation R. K. Friesen et al 2009 ApJ 697 1457 DOI 10.1088/0004-637X/697/2/1457

0004-637X/697/2/1457

ABSTRACT

We present combined interferometer and single dish telescope data of NH3 (J, K) = (1,1) and (2,2) emission toward the clustered star forming Ophiuchus B, C, and F Cores at high spatial resolution (∼1200 AU) using the Australia Telescope Compact Array, the Very Large Array, and the Green Bank Telescope. While the large-scale features of the NH3 (1,1) integrated intensity appear similar to 850 μm continuum emission maps of the Cores, on 15'' (1800 AU) scales we find significant discrepancies between the dense gas tracers in Oph B, but good correspondence in Oph C and F. Using the clumpfind structure identifying algorithm, we identify 15 NH3 clumps in Oph B, and three each in Oph C and F. Only five of the Oph B NH3 clumps are coincident within 30'' (3600 AU) of a submillimeter clump. We find vLSR varies little across any of the cores, and additionally varies by only ∼1.5 km s−1 between them. The observed NH3 line widths within the Oph B and F Cores are generally large and often mildly supersonic, while Oph C is characterized by narrow line widths which decrease to nearly thermal values. We find several regions of localized narrow line emission (Δv ≲ 0.4 km s−1), some of which are associated with NH3 clumps. We derive the kinetic temperatures of the gas, and find they are remarkably constant across Oph B and F, with a warmer mean value (TK = 15 K) than typically found in isolated regions and consistent with previous results in clustered regions. Oph C, however, has a mean TK = 12 K, decreasing to a minimum TK = 9.4 K toward the submillimeter continuum peak, similar to previous studies of isolated starless clumps. There is no significant difference in temperature toward protostars embedded in the Cores. NH3 column densities, N(NH3), and abundances, X(NH3), are similar to previous work in other nearby molecular clouds. We find evidence for a decrease in X(NH3) with increasing N(H2) in Oph B2 and C, suggesting the NH3 emission may not be tracing well the densest core gas.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

Stars form out of the gravitational collapse of centrally condensed clumps5 of dense molecular gas. Recent years have seen leaps forward in our understanding of the structure and evolution of isolated, star forming clumps. Most star formation, however, occurs in clustered environments (Lada & Lada 2003). These regions are more complex, with complicated observed geometries, and contain clumps which tend to have higher densities and more compact sizes than those found in isolation (Motte et al. 1998; Ward-Thompson et al. 2007). It is likely that due to these differences the star formation process in clustered regions proceeds differently than in the isolated cases. Characterizing the physical and chemical structures of these more complicated regions is thus the first step toward a better understanding of the process of clustered star formation.

It is now clear that molecular clumps become extremely chemically differentiated, as many molecules commonly used for tracing molecular gas, such as CO, become severely depleted in the innermost regions through adsorption onto dust grains (see, e.g., Di Francesco et al. 2007 for a review). An excellent probe, therefore, of dense clump interiors is the ammonia molecule (NH3), with a relatively high critical density (ncr ∼ 104 cm−3 for the (1,1) and (2,2) inversion transitions) and apparent resistance to depletion until extreme densities and low temperatures are reached in a starless core's evolution (Tafalla et al. 2004; Aikawa et al. 2005; Flower et al. 2006). The additional kinematic information provided by line observations are complementary to continuum observations of emission from cold dust, and the ammonia molecule in particular allows the determination of the gas kinetic temperature and density structure due to hyperfine transitions of its metastable states (Ho & Townes 1983).

The nearby Ophiuchus molecular cloud, containing the dark L1688 region, is our closest example of ongoing, clustered star formation. The central Ophiuchus cloud has been surveyed extensively in millimeter (Young et al. 2006; Stanke et al. 2006; Motte et al. 1998) and submillimeter (Johnstone et al. 2004, 2000b) continuum emission. These observations have revealed a highly fragmented complex of star-forming clumps with masses M ∼ 0.2–6 M, the majority of which are embedded within larger, highly fragmented structures, called "Cores" for historical reasons (Loren et al. 1990) and named A through I, which reside only in areas of high extinction (Johnstone et al. 2004; Young et al. 2006; Enoch et al. 2007). The total mass of the distinct (sub)millimeter clumps, ∼40–50 M, makes up only 0.5%–3% of the total ∼2000 M mass of the molecular cloud (Young et al. 2006; Johnstone et al. 2004).

Most recent estimates put the distance to the central L1688 cloud region (often also called "ρ Oph") at 120 pc (Loinard et al. 2008; Lombardi et al. 2008; Knude & Hog 1998), in agreement with some older results (de Geus et al. 1989), but a clear consensus has not yet been reached. Mamajek (2008), for example, find a distance of 135 pc toward the cloud using Hipparchos parallax data, while VLBA observations by Loinard et al. (2008) suggest that the Ophiuchus B Core may be further distant than the rest of the cloud, at 165 pc (we also note that Oph B consists of three subcores, B1, B2, and B3, described further in Section 3). This distance is outside the range in median cloud thickness determined by Lombardi et al. (2008) of 28+29−19 pc, but in agreement with an older result by Chini (1981). The stars used to determine the distance to Oph B may, however, be background stars (Lombardi et al. 2008). In the following, we adopt the 120 pc distance to the entire central Ophiuchus region.

In this work, we discuss the results of high-resolution observations of NH3 (1,1) and (2,2) in the Ophiuchus B, C, and F Cores to study the distribution, kinematics, and abundance patterns of the Cores and associated embedded clumps. We find that although the Cores are embedded in the same physical environment, they present very different physical characteristics. We discuss the observations and the combination of interferometer and single dish data in Section 2. In Section 3, we present the data, and detail the hyperfine line fitting procedure and derivations of kinetic temperature TK, NH3 column N(NH3) and space density n(H2) in Section 4 (see also Appendix A). We discuss the results in Section 5, and summarize our findings in Section 6.

2. OBSERVATIONS AND DATA REDUCTION

Figure 1 shows the central Ophiuchus region in 850 μm continuum emission first mapped with the Submillimetre Common User Bolometer Array (SCUBA) at the James Clerk Maxwell Telescope (JCMT) by Johnstone et al. (2000b) and recently re-reduced and combined with all other SCUBA archive data in the region by Jorgensen et al. (2008), following the description in Kirk et al. (2006). The Oph B, C, and F Cores are labeled, and boxes show the approximate areas we mapped at the Green Bank Telescope (GBT), the Australia Telescope Compact Array (ATCA), and the Very Large Array (VLA). The details of all astronomical observations are described below. Table 1 lists the lines observed and their rest frequencies.

Figure 1.

Figure 1. Central region of the Ophiuchus molecular cloud in 850 μm continuum emission originally mapped at the JCMT by Johnstone et al. (2000b). Color scale units are Jy beam−1, where the beam FWHM ≈15''. The B, C, and F Cores are labeled (Oph B3 is not well detected in 850 μm continuum). Rectangles show areas mapped in NH3 (1,1) and (2,2) emission, as well as C2S (21–10) and HC5N (9–8) at the GBT. Stars indicate locations of protostars identified in the infrared with Spitzer (Enoch et al. 2008). VLA and ATCA pointings were placed to provide Nyquist-sampled mosaicking of the indicated regions, with multiple beam overlap in the areas of bright continuum emission.

Standard image High-resolution image

Table 1. Rest Frequencies of all Observed Spectral Lines

Molecule Transition Rest Frequency (GHz) Source
C2S (21–10) 23.3440330 Pickett et al. (1998)
NH3 (1, 1) 23.694495 Ho & Townes (1983)
NH3 (2, 2) 23.722633 Ho & Townes (1983)
HC5N (9 − 8) 23.963888 Myers et al. (1979)

Download table as:  ASCIITypeset image

2.1. Green Bank Telescope

Single-dish observations of emission from the NH3 (J, K) = (1,1) and (2,2) inversion lines, C2S JN = 21–10 and HC5N J = 9 − 8 in the Ophiuchus Cores were obtained using the 100 m Robert C. Byrd GBT, located near Green Bank, WV, USA. The observations were done in frequency-switching mode, using the GBT K-band (upper) receiver as the front end, and the GBT spectrometer as the back end. This setup allowed the simultaneous observation of all lines in four 50 MHz-wide IFs, each with 8192 spectral channels, giving a frequency resolution of 6.104 kHz, or 0.077 km s−1 at 23.694 GHz.

The data were taken using the GBT's on-the-fly (OTF) mapping mode, using in-band frequency switching with a throw of 4 MHz. In OTF mode, a map is created by having the telescope scan across the target in right ascension (R.A.) at a fixed declination (decl.), or in decl. at a fixed R.A., writing data at a predetermined integration interval. The maps of Oph B1 and B2 were made while scanning only in R.A. at a fixed decl., while for subsequent targets (Oph B3, C, and F) the scanning mode was alternated to avoid artificial striping in the final data cubes. No striping, however, is apparent in the final B1 or B2 images. At the observing frequency of 23 GHz, the telescope beam was approximately 32'' FWHM. Subsequent rows or columns were spaced by 13'' in decl. or R.A. to ensure Nyquist sampling. Scan times were determined to ensure either one or two full maps of the observed region could be made between pointing observations. For all observations, pointing updates were performed on the point source calibrator 1622-254 every 45–60 minutes, with corrections approximately 2''–3''. The average telescope aperture efficiency ηA and main beam efficiency ηmb were 0.59 ± 0.05 and 0.78 ± 0.06 respectively, determined through observations of 3C286 at the start of each shift. The absolute flux accuracy is thus ∼8%. The average elevation of Ophiuchus for all observations was approximately 26°.

System temperatures (Tsys) varied between 48 K and 92 K over the observation dates with an average Tsys ∼ 62 K. Table 2 gives the area mapped in each region and the final rms sensitivity in K (TMB).

Table 2. GBT Observation Details by Region

Core Area Mapped (arcmin × arcmin) rms K (TMB)
Oph B1 3 × 3/4 × 5.5a 0.05/0.07b
Oph B2 5 × 4 0.08
Oph B3 3 × 3 0.07
Oph C 3 × 3/4 × 4a 0.05/0.07b
Oph F 3 × 3 0.06

Notes. aFirst values indicate the original OTF map size. Second values are the final map sizes, increased due to observed extent of emission. bThe first rms value is calculated over the central region where all observations overlap. The second value is calculated over the entire map.

Download table as:  ASCIITypeset image

Initial data reduction and calibration were done using the GBTIDL6 package. Zenith opacity values for each night were obtained using a local weather model, and the measured main beam efficiency was used to convert the data to units of main beam temperature, TMB. The two parts of the in-band frequency switched data were aligned and averaged, weighted by the inverse square of their individual Tsys. The data were then converted to AIPS7 SDFITS format using the GBT local utility idlToSdfits. In AIPS, the data were combined and gridded using the DBCON and SDGRD procedures. Finally, the data cubes were written to FITS files using FITTP.

2.2. Australia Telescope Compact Array

Maps of NH3 (1,1) and (2,2) inversion line emission of the Ophiuchus B1, B2, C, and F Cores were made over two separate observing runs at the ATCA. The ATCA is located near the town of Narrabri in New South Wales, Australia, and consists of six antennas, each 22 m in diameter. Five of the six antennas are movable along the facility's east–west line and small north line, while the sixth antenna is permanently placed along the east–west line at a distance of 6 km from the other antennas. An 8 MHz bandwidth with 1024 channels was used, which provided a spectral resolution of 7.81 kHz (0.1 km s−1 at 23.694 GHz). This configuration enabled in-band frequency switching for the observations. The Oph B Cores were observed over four 9 hr tracks of the array (2004 August 5–8), and the Oph C and F Cores were observed over three 9 hr tracks of the array (2005 May 5–7). Both sets of observations were done with the array in its H168 configuration. This is a compact, hybrid configuration, where three of the movable antennas are placed along the east–west line and two antennas are located on the north spur. Baselines ranged from 61.2 to 184.9 m (∼4.7–14.2 kλ) with five antennas. At 23.7 GHz, these observations provided a primary beam (field of view) of 2'.

Maps were made of the cores using separate pointings spaced by $\Theta _N = (2/\sqrt{3}) \, \lambda \, / \, 2D \sim 1\farcm 1$ at 23.7 GHz for Nyquist sampling on a hexagonal grid. Table 3 gives the number of individual pointings required to cover each Core, the multiple-beam overlap area observed, and the final rms sensitivity achieved toward each Core for both the ATCA and the VLA observations (described below).

Table 3. ATCA and VLA Observation Details by Region

Core Overlap Extent (arcmin × arcmin) ATCA Nmosa rms (mJy beam−1) VLA Nmosa rms (mJy beam−1)
Oph B1 2 × 3 3 20 7b 13
Oph B2 5 × 4 10 20 16c 12
Oph C 4 × 4 7 30 7 10
Oph F 3 × 2 3 30 5 15

Notes. The spectral resolution of the observations is 0.1 km s−1 (ATCA) and 0.3 km s−1 (VLA). aNumber of individual pointings in mosaic observations. bPointings for B1 were aligned and spaced to provide continuous coverage between B1 and B2. cAdditional pointings added to the B2 Core to include B3 in the overlap region.

Download table as:  ASCIITypeset image

Observations cycled through each individual pointing between phase calibrator observations to maximize uv-coverage and minimize phase errors. The phase calibrator, 1622-297, was observed every 20 minutes, and was also used to check pointing every hour. Flux and bandpass calibration observations were performed every shift on the bright continuum sources 1253-055, 1934-638, and 1921-293.

The ATCA data were reduced using the MIRIAD data reduction package (Sault et al. 1995). The data were first flagged to remove target observations unbracketed by phase calibrator measurements, data with poor phase stability or anomalous amplitude measurements. The majority of the data were good, as the weather was stable during the observations. Much of the data from baselines involving the 6 km antenna, however, were flagged due to poor phase stability. The bandpass, gains, and phase calibrations were applied, and the data were then jointly deconvolved. First, the data were transformed from the spatial frequency (u, v)-plane into the image plane using the task INVERT and natural weighting to maximize signal to noise. The data were then deconvolved and restored using the tasks MOSSDI and RESTOR to remove the beam response from the image. MOSSDI uses a Steer–Dewdney–Ito CLEAN algorithm (Steer et al. 1984). The cleaning limit was set at twice the rms noise level in the beam overlap region for each object. Clean boxes were used to avoid cleaning noise in the outer regions with less beam overlap. Applying natural weighting provided a final synthesized beam of ∼8'' × 10'' FWHM.

2.3. Very Large Array

Maps of NH3 (1,1) and (2,2) emission were made at the VLA near Socorro, NM, USA over the period of 2007 January 20 through 2007 February 11. Nine observing shifts were allotted to the project in the array's DnC configuration, each 5 hr in duration covering the LST range 14:00–19:00. The DnC configuration is a hybrid of the most compact D configuration with the next most compact C configuration. This setup ensures a more circular beam shape for southern sources like Ophiuchus while retaining the sensitivity of the D configuration.

For these observations, we used a correlator setup with two IFs, each with a bandwidth of 3.125 MHz with 24.414 kHz spectral resolution (0.3 km s−1). While not providing the same high spectral resolution obtained at the ATCA, this setup enabled simultaneous observations of the (1,1) and (2,2) lines, and allowed the main hyperfine component and the two middle satellite components of the (1,1) line to be contained within the band. Mosaic maps were made with Nyquist-spaced (∼1farcm0 at 23 GHz) individual pointings. Observations cycled through pointings between phase calibrator observations. Table 3 gives the pointing and rms sensitivity information for each Core.

During these observations, several antennas in the array had been upgraded as part of the Expanded VLA (EVLA) project. Online Doppler tracking at this time was not yet available, and the observations were thus obtained in a fixed frequency mode, with line sky frequencies calculated using the NRAO's online Dopset tool. The observing frequencies were updated frequently, with phase calibrator observations on 1625-254 before and after a frequency change to avoid phase jump problems. Bandpass and absolute flux calibration were done for each shift using observations of 1331+305.

The data were checked, flagged, and calibrated using the NRAO Astronomical Image Processing System (AIPS), following the procedures outlined in the AIPS Cookbook.8 In addition, special processing was required to account for differences in bandpass shape and antenna sensitivity between the VLA and EVLA antennas. System temperatures were lacking for EVLA antennas, so the VLA back-end Tsys values were used when reading the data into an AIPS uv-database. A bandpass table was then created from the line data set and applied to a spectrally averaged "channel 0" data set. The normal VLA calibration steps were then followed. The calibrated uv files were then written to FITS format and imported to MIRIAD, where the data were deconvolved and restored. Applying natural weighting and a taper of 8'' × 6'' provided a final synthesized beam of ∼10farcs5 × 8farcs5 FWHM.

2.4. Combining Single Dish and Interferometer Data Sets

Since none of the antennas in the ATCA and the VLA can act as single radio dishes, there is necessarily an upper limit to the size of structure to which each is sensitive. This upper scale limit is dependent on the shortest spacing between two antennas in the array, and the missing information is thus referred to as the short- or zero-spacing problem. Mosaicking helps to recover short spacing information. For complex sources with emission on many scales, however, combining interferometer data with single-dish observations provides more complete coverage of the u-v plane and thus creates a more accurate representation of the true source emission structure. Ideally, the single dish diameter should be larger than the minimum interferometer baseline to ensure maximal overlap in the uv-plane and determine accurately flux calibration factors between the observations.

Data from each interferometer were first combined separately with the single dish observations. The GBT data were regridded to match the interferometer data in pixel scale and spectral resolution. The data were then converted to units of Jy beam−1 for combination with the interferometer data using the average beam FWHM measured at the GBT during the observations. Combination of the data was done using MIRIAD's IMMERGE task. IMMERGE combines deconvolved interferometer data with single-dish observations by Fourier transforming both data sets and combining them in the Fourier domain, applying tapering functions such that at small spacings (low spatial frequencies) the single-dish data are more highly weighted than the interferometer data, while conversely at high spatial frequencies the interferometer data are weighted more highly. The flux calibration factor between the two data sets was calculated in IMMERGE by specifying the overlapping spatial frequencies between the GBT and the interferometers. We took the overlap region in both cases to be 35–100 m (2.7–7.7 kλ), yielding a flux calibration factor of ∼1.4 between the GBT and ATCA data sets, and ∼1.0 between the GBT and VLA data sets in the NH3 (1,1) line emission. These factors were then applied to the NH3 (2,2) data sets. The final resolution of the combined data is the same as that of the interferometer data.

To combine all three data sets, the ATCA and VLA data were imaged together using the INVERT task, applying natural weighting and taper as described for the VLA imaging. The interferometer data were then cleaned and combined with the GBT data as described above. The overlap region was taken to be 35–100 m, yielding a flux calibration factor of ∼1.3. By convolving the combined data to match the 32'' resolution of the GBT data, we estimate the total flux of the combined image recovers nearly all (∼98%) the flux in the single-dish image. These data were used to identify structures in the NH3 data cubes using an automated structure finding algorithm (see Section 3). The data were also tapered to a slightly lower resolution of 15'' FWHM to provide higher signal-to-noise ratios (S/Ns) for a multiple component hyperfine line fitting routine.

3. RESULTS

3.1. Comparison with Submillimeter Dust Continuum Emission

We first discuss the NH3 (1,1) intensity in the combined data sets, and compare the distribution of NH3 emission and 850 μm continuum emission in the Oph Cores as shown in Figure 1. Figures 2(a), 3(a), and 4(a) show the combined NH3 (1,1) line emission at ∼10farcs6 × 8farcs5 FWHM resolution. The emission has been integrated over the central hyperfine components in the Oph B, C, and F Cores, respectively, with a clip of ∼2× the map rms noise level. (Since the outer edge of the combined maps have higher rms noise levels, integrating only over, i.e., the "main component" of NH3 reduced the amount of signal included in the noisy outer regions.) The respective 850 μm emission for each core at ∼15'' FWHM resolution, i.e., lower than the resolution of the combined NH3 data, is shown in Figures 2(b), 3(b), and 4(b). We also show locations of submillimeter clumps identified with the 2D version of clumpfind (Jorgensen et al. 2008). In Figure 2(b), we additionally label the ∼1 M continuum object B2-MM8 (Motte et al. 1998).

Figure 2.

Figure 2. Observations of the Oph B Core (including B1, B2, and B3). At lower left in each panel, ovals show the respective resolutions of the data. (a) Integrated NH3 (1,1) main component intensity obtained with the GBT, ATCA, and VLA telescopes with a synthesized beam FWHM 10farcs5 × 8farcs5. Emission was integrated over spectral channels with intensities ⩾±2σ rms. The color scale is in K km s−1 (TB). Contours begin at 3 K km s−1 and increase by 1.5 K km s−1. In both plots, stars denote locations of Class I protostars (Enoch et al. 2008), while triangles indicate the positions of NH3 clumps as identified by clumpfind. (b) Continuum emission at 850 μm in Oph B at 15'' resolution as in Jorgensen et al. (2008). The color scale is in mJy beam−1. Contours begin at 50 mJy beam−1 and increase by 100 mJy beam−1. Squares show locations of submillimeter clumps (Jorgensen et al. 2008). We also label the millimeter clump B2-MM8 (Motte et al. 1998) and protostars identified previous to Enoch et al. (2008).

Standard image High-resolution image
Figure 3.

Figure 3. Observations of the Oph C Core. At lower left in each panel, ovals show the respective resolutions of the data. (a) Integrated NH3 (1,1) main component intensity obtained with the GBT, ATCA, and VLA telescopes, summed over spectral channels with intensities ⩾2σ rms. The color scale is in units of K km s−1 (TB). Contours begin at 3 K km s−1 (TB) and increase by 1.5 K km s−1. In both plots, triangles indicate the positions of NH3 clumps as identified by clumpfind. (b) Continuum emission at 850 μm in Oph C at 15'' resolution as in Jorgensen et al. (2008). Contours and color range as in Figure 2. Squares show locations of submillimeter clumps (Jorgensen et al. 2008).

Standard image High-resolution image
Figure 4.

Figure 4. Observations of the Oph F Core. At lower left in each panel, ovals show the respective resolutions of the data. (a) Integrated intensity NH3 (1,1) main component intensity obtained with the GBT, ATCA, and VLA telescopes, summed over spectral channels with intensities ⩾2σ rms. The color scale is in units of K km s−1 (TB). Contours begin at 3 K km s−1 (TB) and increase by 1 K km s−1. In both plots, stars denote locations of Class I protostars (Enoch et al. 2008), while triangles indicate the positions of NH3 clumps as identified by clumpfind. (b) Continuum emission at 850 μm in Oph F at 15'' resolution as in Jorgensen et al. (2008). Contours and color range as in Figure 2. Squares show locations of submillimeter clumps (Jorgensen et al. 2008). We also label protostars identified previous to Enoch et al. (2008).

Standard image High-resolution image

Figures 23, and 4 also show locations and labels of "cold" young stellar objects (YSOs; based on bolometric temperatures derived from fitting their spectral energy distributions) detected and classified through Spitzer infrared observations (Enoch et al. 2008) overlaid on the submillimeter and NH3 emission. The objects plotted were all identified as Class I protostars (no Class 0 protostars have been associated with Oph B, C, or F). Oph B2 is associated with three YSOs. Of these, two are previously known (IRS45/GY273 and IRS47/GY279). Based on association with a continuum emission object (Enoch et al. 2008) and observed infrared colors (Jorgensen et al. 2008), either three or two protostars in Oph B2 are embedded in gas and dust. One additional Class I protostar is located between B1 and B2, and is only associated with diffuse NH3 emission. Oph B1 and B3 appear starless. No embedded protostars are associated with Oph C by Enoch et al. (2008), but a "Candidate YSO" is identified by Jorgensen et al. (2008) 30'' south of the core continuum peak (R.A. 16:26:59.1, decl. −24:35:03). Based on the different classifications by the two papers, the significant offset from the continuum emission peak, and the lack of any clear influence on the gas in our data, we will discuss Oph C assuming it is not associated with a deeply embedded object. Four protostars are associated with Oph F, all of which have been previously identified (see Figure 4 for object names). Three are likely embedded in the Core.

Peaks of NH3 integrated intensity can be used to surmise the presence of "objects," but such identifications can ignore any differences in velocity between adjacent dense gas. We therefore used the 3D version of the automated structure-finding routine clumpfind (Williams et al. 1994) to identify distinct NH3 emission objects in the combined NH3 (1,1) data cube, which we will henceforth call "NH3 clumps." clumpfind uses specified brightness contour intervals to search through the data cube for distinct objects identified by closed contours. The size of the identified clumps are determined by including adjacent pixels down to an intensity threshold, or until the outer edges of two separate clumps meet. The clump location is defined as that of the emission peak. clumpfind was used only on the main emission component, where multiple hyperfine components are sufficiently blended to present effectively a single line given the 0.3 km s−1 spectral resolution and typically wide line widths found. The standard interval of 2 times the data rms between contours, with a slightly larger threshold value worked well to separate distinct emission components in all clumps. A lower limit of 1 K with intervals of 0.4 K, ∼2 times the cube rms, identified separate emission peaks sufficiently for Oph B and C. Oph F was sufficiently fit using a lower limit of 1.2 K and intervals of 0.65 K, as the rms of the combined data was slightly higher. Even so, some identified clumps appeared by eye to be noise spikes at the map edges, and these were culled from the final list, as well as any clumps which contained fewer pixels than in the synthesized beam. Table 4 lists the locations, FWHMs, effective radii, and peak brightness temperatures for clumpfind-identified clumps in Oph B, C, and F, and the clump locations are overlaid on Figures 24. The centroid velocities and line widths of the NH3 clumps were determined through spectral line fitting (described further in Section 4), which provided more accurate measures of vLSR and Δv given the hyperfine structure of the NH3 lines.

Table 4. NH3 (1,1) Clumpfind Peaks and Parameters

ID R.A. J2000 Decl. J2000 FWHMx× FWHMya (AU) Peak (K)
B1-A1 16 27 9.7 −24 27 43.0 3900 × 6700 4.90
B1-A2 16 27 12.4 −24 29 49.0 6800 × 7800 5.10
B1-A3 16 27 15.4 −24 30 29.1 5300 × 5000 5.60
B1-A4 16 27 16.7 −24 30 59.1 3700 × 2900 4.90
B2-A1 16 27 13.8 −24 27 35.0 2300 × 2500 5.20
B2-A2 16 27 16.0 −24 27 39.0 5600 × 6100 5.10
B2-A3 16 27 19.0 −24 27 11.0 4700 × 6700 3.90
B2-A4 16 27 24.2 −24 27 29.0 6400 × 6000 5.10
B2-A5 16 27 25.0 −24 26 55.0 5300 × 4700 5.00
B2-A6 16 27 29.0 −24 26 59.0 3800 × 6700 5.70
B2-A7 16 27 32.5 −24 26 57.0 3800 × 5200 6.70
B2-A8 16 27 32.8 −24 26 17.0 3600 × 3600 4.00
B2-A9 16 27 33.5 −24 27 43.0 5000 × 3600 3.60
B2-A10 16 27 33.8 −24 26 14.9 2700 × 3700 4.00
B3-A1 16 27 21.8 −24 24 42.9 3900 × 6100 3.60
C-A1 16 26 59.0 −24 34 14.8 5000 × 5600 5.40
C-A2 16 26 59.1 −24 32 54.8 4500 × 4400 4.30
C-A3 16 27 1.2 −24 34 34.8 6000 × 5400 5.50
F-A1 16 27 21.9 −24 39 52.0 3600 × 3200 4.50
F-A2 16 27 24.2 −24 40 52.1 4400 × 5900 4.90
F-A3 16 27 27.1 −24 40 58.1 2400 × 5000 5.50

Note. aFWHM values calculated by clumpfind have not been deconvolved with the synthesized beam of the combined data.

Download table as:  ASCIITypeset image

In Oph B, Figure 2 shows that the large-scale structure of the Core is similar in both line and continuum emission, but some significant differences are also apparent. For example, the integrated line emission displays a less pronounced division between the B1 and B2 Cores seen in the continuum, and indeed shows significant filamentary structure in the region connecting B1 and B2 where little to no continuum emission is observed. Similarly, strong line emission is present in the southeastern edge of B2 where the continuum map shows relatively little emission. Additionally, while B2 is the stronger continuum emitter, B1 is significantly brighter in integrated line emission. Strong NH3 emission in the northern half of B1 extends beyond the continuum contours, and becomes significantly offset from the bulk of the continuum emission to the east. The NH3 (1,1) observations of the Oph B Core region also reveal a narrow-line emission peak north of the western edge of B2, which is coincident with a DCO+ object, Oph B3 (Loren et al. 1990). B3 is just visible at the lowest contour in the NH3 integrated intensity map but is not visible in the continuum map.

Furthermore, although the peaks of continuum emission and line emission are often located in the same vicinity in B1 and B2, the brightest continuum peaks and the integrated line emission maxima are typically noncoincident. Overall, the mean separation between an NH3 integrated intensity peak and the nearest 850 μm continuum peak in B1 and B2 is ∼22'' (2600 AU), or ∼2 times the NH3 FWHM resolution. Fifteen clumpfind-identified NH3 clumps are identified in the combined Oph B map, with an average minimum separation between NH3 clumps of 47'' (5600 AU), or 40'' (4800 AU), if Oph B3 is not included. The mean minimum distance between NH3 clumps and submillimeter continuum clumps is 44'' (5300 AU), or ∼4 times the NH3 FWHM resolution. Only five of the fifteen NH3 clumps are located within 30'' (3600 AU) of a submillimeter continuum clump.

No protostars in Oph B are found at positions of NH3 (1,1) integrated intensity maxima nor are they associated with identified NH3 clumps. One protostar, IRS47/GY279, is located south of the NH3 clump we identify as B2-A6. The offset between the NH3 clump peak and the protostar, however, is ∼30'', or approximately 3 times the angular resolution of the combined NH3 data. A second protostar, IRS45/GY273, is coincident with a submillimeter continuum peak but has little associated NH3 emission. Either of these protostars may be the source of an east–west aligned outflow recently proposed by Kamazaki et al. (2003) from CO observations of B2. The third, previously unidentified protostar, seen west of NH3 clump B2-A4, is located within a narrow (∼30'') NH3 integrated intensity minimum between the B2 Core and the filament connecting Oph B1 and B2.

Figure 3 shows that Oph C has more similar overall structure when traced by the continuum and integrated NH3 (1,1) line emission than Oph B. Extended emission in Oph C is elongated along a southeast-northwest axis and contains a single integrated intensity peak. A thin (∼30'') filament of faint emission extends to the north of the central peak. The submillimeter continuum emission is largely coincident with the integrated NH3 contours, but the continuum emission peak is offset to the NH3 integrated intensity peak by ∼20''. clumpfind separates the central NH3 emission into two clumps, C-A1 and C-A3, and finds a third object, C-A2, at the tip of the northern emission extension. C-A1 and C-A3 are found to the northwest and southeast (30'' offset and 15'' offset, respectively) of the centers of both the NH3 integrated intensity and of the continuum emission. Continuum emission also extends in the direction of the faint NH3 northern extension, but there is no secondary peak present.

Like Oph C (but unlike Oph B), Figure 4 shows that Oph F also has very similar structure when traced by either the submillimeter continuum emission or integrated NH3 intensity. Unlike both Oph B and C, the clumpfind-identified NH3 clumps are coincident with the integrated intensity peaks. F-A3 is nearly coincident (within a beam FWHM) with a submillimeter continuum clump and is additionally coincident with an embedded protostar, IRS43/GY265. F-A2 is associated with a second embedded protostar (CRBR65) and continuum emission, but not an identified continuum clump. A thin filament (∼15'') extends to the northwest and the third NH3 clump, F-A1, which is also coincident with extended continuum emission but no identified clump. A third embedded protostar (IRS44/GY259) in the northeast is coincident with a continuum peak, but has no associated NH3 emission. A fourth protostar, in the south-west, may be coincident with some unresolved NH3 emission, but is located in a section of the map with larger rms values and consequently the small integrated intensity peak seen at that location may be simply noise.

The discrepancies between NH3 and submillimeter continuum emission in Oph B are in contrast to earlier findings of extremely high spatial correlation between the two gas tracers in isolated, low-mass starless clumps. For example, Tafalla et al. (2002) found that both the millimetre continuum and integrated NH3 (1,1) and (2,2) line intensity were compact and centrally concentrated in a survey of five starless clumps, including L1544 in the Taurus molecular cloud. In these clumps, the integrated intensity maxima of both transitions are approximately coincident (within the 40'' angular resolution of the NH3 observations) with the continuum emission peaks. This same coincidence between NH3 and millimeter continuum was found in B68 (Lai et al. 2003). When observed at higher angular resolution, NH3 emission in L1544 remained coincident with the continuum but the line integrated intensity peak was offset by ∼20''. This offset was explained, however, as being due to the NH3 emission becoming optically thick (Crapsi et al. 2007).

It is also possible that different methods of identifying structure in molecular gas (such as the gaussclump method of Stutzki & Guesten 1990, or using dendrograms as in Rosolowsky et al. 2008b, for example) would create a different "clump" list than presented here. We have additionally compared our results in Oph B with the locations of millimeter objects identified using multiwavelet analysis by Motte et al. (1998), and find a smaller yet still significant mean minimum distance of 27'' between NH3 clumps and millimeter objects. Given the severe positional offsets in some locations between the NH3 emission and submillimeter continuum, it is unlikely that different structure-finding methods would provide substantially different results.

3.2. Single Dish C2S and HC5N Detections

The GBT observations of C2S (21–10) emission only resulted in single, localized detections in Oph B1 and Oph C, while only Oph C had a single, localized detection in HC5N (9–8). The C2S emission in B1 was confined to a single peak at its southern tip. The single-dish spectra of all observed species at the C2S peak locations in B1 and C are presented in Figure 5 (note that in Figure 5, the NH3 line is so narrow in Oph C that we are detecting the hyperfine structure of the (1,1) line). The integrated intensity GBT maps of all molecules observed in Oph C are shown in Figure 6. Within the ∼30'' resolution limits of the GBT data, the NH3, C2S and HC5N spectral line integrated intensity peaks overlap with the local 850 μm continuum emission peak in Oph C. We fit the spectra of the two C2S and single HC5N detections with single Gaussians to determine their respective vLSR, line width Δv, and peak intensity of the line in TMB units. We additionally fit a 2D Gaussian to the integrated intensity maps to determine the FWHMs of the emitting regions. We find a beam-deconvolved FWHM = 9200 AU and 5800 AU for the C2S and HC5N emission, respectively, in Oph C. The C2S emission in Oph B1 is elongated, with a FWHM = 9400 AU in R.A. but only 3700 AU in decl. for an effective FWHM = 5100 AU. The results of the Gaussian fitting are listed in Table 5. See Section 4.4.2 for further analysis of C2S and HC5N.

Figure 5.

Figure 5. (a) Spectra of all species observed at the GBT at the integrated intensity C2S (21–10) peak in Oph B1. The NH3 (1,1) and (2,2) and C2S baselines are offset from 0 for clarity. (b) Spectra of all species observed at the GBT at the integrated intensity C2S (21–10) peak in Oph C. The NH3 (1,1) and (2,2) and C2S baselines are offset for clarity.

Standard image High-resolution image
Figure 6.

Figure 6. GBT observations of Oph C. The circle in each panel gives the relative FWHM of the GBT data. (a) The Ophiuchus C Core in 850 μm continuum emission at 15'' resolution (grayscale, Jy beam−1 units) overlaid with continuum contours (gray). Black contours show the integrated NH3 (1,1) intensity at ∼32'' resolution, beginning at 3 K km s−1 and increasing by 1.5 K km s−1. For all plots, emission was integrated over spectral channels with intensities ⩾±2σ rms. (b) 850 μm continuum emission in Oph C overlaid with black NH3 (2,2) integrated emission contours. Contours begin at 0.3 K km s−1 and increase by 0.1 K km s−1. (c) 850 μm continuum emission in Oph C overlaid with black C2S (21–10) integrated emission contours. Contours are 0.2, 0.25, 0.3 K km s−1. (d) 850 μm continuum emission in Oph C overlaid with black HC5N (9 − 8) integrated emission. Contours are 0.1, 0.2 K km s−1.

Standard image High-resolution image

Table 5. GBT C2S (21–10) and HC5N (9–8) Peak Parameters

ID R.A. J2000 Decl. J2000 FWHM (AU) Vlsr (km s−1) Δv (km s−1) TMB (K) N (1012 cm−2) X (10−10)
B1-C2S 16 27 13.1 −24 30 50 5100 3.67(2) 0.33(5) 0.36(3) 4.3(7) 3.1(7)
C-C2S 16 26 59.8 −24 34 13 9200 3.95(1) 0.33(2) 1.12(4) 10.6(8) 1.5(1)
C-HC5N 16 26 58.8 −24 34 30 5800 3.94(1) 0.35(2) 0.68(3) 4.0(3) 0.46(4)

Download table as:  ASCIITypeset image

4. NH3 LINE ANALYSIS

4.1. NH3 Hyperfine Structure Fitting

The metastable J = K rotational states of the symmetric-top NH3 molecule are split into inversion doublets due to the ability of the N atom to quantum tunnel through the hydrogen atom plane. Quadrupole and nuclear hyperfine effects further split these inversion transitions, resulting in hyperfine structure (HFS) of the (J, K) = (1,1) transition, for example, containing 18 separate components. This hyperfine structure allows the direct determination of the optical depth of the line through the relative peaks of the components. Additionally, since transitions between K-ladders are forbidden radiatively, the rotational temperature describing the relative populations of two rotational states, such as the (1,1) and (2,2) transitions, can be used to determine directly the kinetic gas temperature (Ho & Townes 1983).

For a given (J, K) inversion transition of NH3 in local thermodynamic equilibrium (LTE), the observed brightness temperature T*A as a function of frequency ν can be written as

Equation (1)

assuming the excitation conditions of all hyperfine components are equal and constant (i.e., Tex = Tex). Here, ηMB is the main beam efficiency, Φ is the beam filling factor of the emitting source, Tex is the line excitation temperature, Tbg = 2.73 K is the temperature of the cosmic microwave background, and J(T) = (hν/k)[exp(hν/kT) − 1]−1. The line opacity as a function of frequency, τν, is given by

Equation (2)

where N is the total number of hyperfine components of the (J, K) transition (N = 18 for the (1,1) transition and N = 21 for the (2,2) transition). For a given jth hyperfine component, aj is the emitted line fraction and νj is the expected emission frequency. The observed frequency of the brightest line component is given by ν0, with a FWHM Δν. Values of aj and νj were taken from Kukolich (1967). Here, we assume Φ = 1. If the observed emission does not entirely fill the beam, the determined Tex will be a lower limit. In regions where the emission is very optically thin (τ < <1) there is a degeneracy between τ and Tex and solving for the parameters independently becomes impossible. We restricted our analysis to regions where the NH3 (1,1) intensity in the central component is greater than 2 K, which corresponds roughly in our data to an S/N of 8 − 10 in the main component and ∼2–3 in the satellite components. With this restriction, we also find τ ≳ 0.5 throughout the regions discussed.

To improve the S/N of the data and to match the resolution of the 850 μm continuum data, we first convolved the combined data to a final FWHM of 15'' (from 10farcs6 × 8farcs5), and then binned the convolved data to 15'' × 15'' pixels. Assuming Gaussian profiles, the 18 components of the NH3 (1,1) emission line were fit simultaneously using a chi-square reduction routine custom written in idl. The returned fits provide estimates of the local standard of rest (LSR) line centroid velocity (vlsr), the observed line FWHM (Δvobs), the opacity of the line summed over the 18 components (τ), and [J(Tex) − J(Tbg)]. The satellite components of the (2,2) line are not visible above the rms noise of our data. These data were consequently fit (again in idl) with a single Gaussian component.

The line widths determined by the hyperfine structure fitting routine are artificially broadened by the velocity resolution (0.3 km s−1) of the observations. To remove this effect, we subtract in quadrature the resolution width, Δvres, from the observed line width, Δvobs, such that $\Delta v_{\rm line} = \sqrt{\Delta v_{\rm obs}^2 - \Delta v_{\rm res}^2}$. In the following, we simply use Δv = Δvline for clarity. The limitations of the moderately poor velocity resolution are discussed further in Appendix B, but do not significantly impact our analysis. In regions where lines are intrinsically narrow, such as Oph C and parts of Oph F, the derived line widths may be overestimated by up to ∼20%–35%.

The uncertainties reported in the returned parameters are those determined by the fitting routine, and do not take the calibration uncertainty of ∼8% into account. The calibration uncertainty affects neither the derived parameters that are dependent on ratios of line intensities, such as the opacity and kinetic temperature TK, nor the uncertainties returned for vLSR or Δv. The excitation temperature, however, as well as the column densities and fractional NH3 abundances discussed below (see Section 4.4) are dependent on the amplitude of the line emission, and are thus affected by the absolute calibration uncertainty.

Table 6 lists the mean, rms, minimum, and maximum values of vlsr, Δv, τ, and Tex found in each of the Cores using the above restrictions for the combined NH3 line emission. In the following sections, we describe in detail the results of the line fitting and examine the resulting line centroid velocities and widths, as well as Tex and τ. In addition, we use the fit parameters to calculate the gas kinetic temperature (TK), nonthermal line widths (σNT), NH3 column density (N(NH3)), gas density (n(H2)) and NH3 abundance (X(NH3)) across all the Cores, as described further in Sections 4.2, 4.3, and 4.4. Table 7 summarizes the mean, rms, minimum, and maximum values of TK, σNT, N(NH3), n(H2), and X(NH3) for each core. For each NH3 clump, Table 8 summarizes the mean, rms, minimum, and maximum values of all determined parameters (means were obtained by uniformly weighting each pixel).

Table 6. NH3 (1,1) Line Characteristics in Combined Data

Core Value Mean RMS Min Max
Oph B vlsr (km s−1) 3.96 0.24 3.15 4.57
  Δv (km s−1) 0.83 0.21 0.08 1.37
  τ 1.6 0.8 0.5 4.7
  Tex (K) 9.9 2.3 5.4 19.5
Oph C vlsr (km s−1) 4.01 0.07 3.81 4.10
  Δv (km s−1) 0.37 0.13 0.11 0.70
  τ 4.7 2.6 0.7 10.8
  Tex (K) 6.9 1.2 5.1 11.6
Oph F vlsr (km s−1) 4.34 0.12 4.24 4.82
  Δv (km s−1) 0.53 0.30 0.10 1.13
  τ 1.9 0.8 0.7 3.6
  Tex (K) 9.5 2.5 5.0 14.3

Download table as:  ASCIITypeset image

Table 7. Physical Properties of Filaments Derived From Fitted Parameters

Core Value Mean RMS Min Max
Oph B Tk (K) 15.1 1.8 12.3 22.9
  σNT (km s−1) 0.35 0.09 0.04 0.57
  σNT/cs 1.5 0.4 0.5 2.5
  N(NH3) (× 1014 cm−2) 4.6 2.1 0.7 12
  n(H2) (× 104 cm−3) 5.9 10 0.8 77
  X(NH3) (× 10−8) 2.9 2.1 0.5 8.7
Oph C Tk (K) 12.0 1.6 9.4 19.8
  σNT (km s−1) 0.14 0.06 0.0 0.29
  σNT/cs 0.6 0.2 0.0 1.2
  N(NH3) (× 1014 cm−2) 5.3 2.4 1.6 13
  n(H2) (× 104 cm−3) 2.0 1.0 0.8 4.2
  X(NH3) (× 10−8) 1.7 0.6 0.6 3.1
Oph F Tk (K) 15.5 2.5 12.8 22.9
  σNT (km s−1) 0.21 0.13 0.05 0.47
  σNT / cs 0.9 0.6 0.2 2.2
  N(NH3) (× 1014 cm−2) 2.6 1.6 0.8 6.6
  n(H2) (× 104 cm−3) 3.6 2.4 0.9 10.6
  X(NH3) (× 10−8) 1.8 0.6 0.5 3.1

Download table as:  ASCIITypeset image

4.2. Line Centroids and Widths

Figures 7(a), 8(a), and 9(a) show maps of vLSR of the fitted NH3 (1,1) line in Oph B, C, and F respectively from the combined, smoothed and regridded data. These maps reveal that although variations of vLSR are seen within the Cores, they are not that kinematically distinct from each other. For example, only 0.35 km s−1 (≲ the mean Δv) separates the average line-of-sight velocity in Oph B from Oph F. This result agrees with the one-dimensional velocity dispersion of ∼0.36 km s−1 found by André et al. (2007) through N2H+ observations of the Oph Cores.

Figure 7.

Figure 7. (a) Line velocity or vLSR in Oph B. Color scale is in km s−1. The values shown are those derived after convolving the data to 15'' resolution and then regridding into 15'' × 15'' pixels. In all plots, contours show integrated NH3 (1,1) intensity convolved to 15'' resolution, beginning at 4.5 K km s−1 and increasing by 1.5 K km s−1. Stars indicate protostar positions and triangles indicate positions of NH3 clumps. (b) Fitted Δv in Oph B. Color scale is in km s−1. (c) Tk in Oph B. Grayscale from 10 K to 20 K. (d) Ratio of the nonthermal to thermal line width components. Grayscale from σNT/cs = 0 to 2.4. (e) Total column density of H2 derived from 850 μm dust continuum observations in Oph B, regridded to match the combined NH3 observations. The dust temperature Td per pixel was assumed to be equal to the gas temperature Tk derived from HFS line fitting of the NH3 observations. The N(H2) values shown have been divided by 1021. (f) Fractional abundance of NH3 divided by 10−9.

Standard image High-resolution image
Figure 8.

Figure 8. (a) Line velocity or vLSR in Oph C. Color scale is in km s−1. The values shown are those derived after convolving the data to 15'' resolution and then regridding into 15'' × 15'' pixels. In all plots, contours show integrated NH3 (1,1) intensity convolved to 15'' resolution, beginning at 4.5 K km s−1 and increasing by 1.5 K km s−1. Stars indicate protostar positions and triangles indicate positions of NH3 clumps. (b) Fitted Δv in Oph C. Color scale is in km s−1. (c) Tk in Oph C. Grayscale from 10 K to 20 K. (d) Ratio of the nonthermal to thermal line width components. Grayscale from σNT/cs = 0 to 2.4. (e) Total column density of H2 derived from 850 μm dust continuum observations in Oph C, convolved to 15'' resolution and regridded to match the combined NH3 observations. The dust temperature Td per pixel was assumed to be equal to the gas temperature TK derived from HFS line fitting of the NH3 observations. The N(H2) values shown have been divided by 1021. (f) Fractional abundance of NH3 divided by 10−9.

Standard image High-resolution image
Figure 9.

Figure 9. (a) Line velocity or vLSR in Oph F. Color scale is in km s−1. The values shown are those derived after convolving the data to 15'' resolution and then regridding into 15'' × 15'' pixels. In all plots, contours show integrated NH3 (1,1) intensity convolved to 15'' resolution, beginning at 3 K km s−1 and increasing by 1 K km s−1. Stars indicate protostar positions and triangles indicate positions of NH3 clumps. (b) Fitted Δv in Oph F. Color scale is in km s−1. (c) Tk in Oph C. Grayscale from 10 K to 20 K. (d) Ratio of the nonthermal to thermal line width components. Grayscale from σNT/cs = 0 to 2.4. (e) Total column density of H2 derived from 850 μm dust continuum observations in Oph F, convolved to 15'' resolution and regridded to match the combined NH3 observations. The dust temperature Td per pixel was assumed to be equal to the gas temperature Tk derived from HFS line fitting of the NH3 observations. The N(H2) values shown have been divided by 1021. (f) Fractional abundance of NH3 divided by 10−9.

Standard image High-resolution image

Table 8. Derived Parameters at NH3 (1,1) Peak Locations

ID vlsr (km s−1) Δv (km s−1) τ Tk (K) Tex (K) σNT (km s−1) N(H2) (1021 cm−2) $N(\mbox{NH}_3)_{\rm {tot}}$ (1014 cm−2) X(NH3) (10−8) n(H2) (104)
B1-A1 4.01(3) 0.85(7) 1.6(3) 16.3(1.9) 9.4(1) 0.35(3)  ⋅⋅⋅  5.0(9)  ⋅⋅⋅  2.8(1)
B1-A2 3.84(2) 0.92(5) 1.5(2) 16.4(1.3) 12.1(1) 0.38(2) 14(3) 6.3(9) 6(1) 6.5(2)
B1-A3 3.98(1) 0.62(3) 3.4(2) 13.7(1.0) 9.7(1) 0.25(2) 18(4) 9.1(7) 7(1) 4.2(1)
B1-A4 3.83(2) 0.54(3) 4.2(2) 12.4(0.9) 8.4(1) 0.22(2) 12(2) 9.5(8) 11(2) 3.2(1)
B2-A1 4.18(2) 0.57(5) 1.6(3) 13.3(1.2) 11.6(1) 0.23(2)  ⋅⋅⋅  4.9(9)  ⋅⋅⋅  12.6(2)
B2-A2 4.13(2) 0.86(5) 0.3(*) 14.1(*) 31.0(*) 0.35(*)  ⋅⋅⋅  3.9(*)  ⋅⋅⋅  0.0(*)
B2-A3 4.10(3) 0.80(7) 1.3(3) 14.2(1.2) 10.2(1) 0.33(3) 15(3) 4(1) 4.2(8) 4.7(1)
B2-A4 4.00(2) 1.00(6) 1.4(2) 13.5(1.0) 11.6(1) 0.42(3) 32(6) 7(1) 3.2(6) 11.7(8)
B2-A5 3.91(2) 0.79(5) 2.4(2) 13.2(1.0) 9.7(1) 0.32(2) 45(9) 8.6(9) 2.7(5) 4.8(2)
B2-A6 4.32(2) 0.66(4) 1.3(2) 14.5(1.1) 14.1(1) 0.27(2) 33(7) 5.2(9) 2.2(4) 72(26)
B2-A7 4.57(1) 0.33(2) 2.8(2) 13.9(1.1) 11.9(1) 0.12(1) 35(7) 4.8(6) 2.0(4) 11.5(8)
B2-A8 4.35(3) 0.95(8) 0.5(*) 14.6(*) 16.9(1) 0.39(*) 43(8) 3.5(*) 1.2(2) 0.0(*)
B2-A9  ⋅⋅⋅   ⋅⋅⋅   ⋅⋅⋅   ⋅⋅⋅   ⋅⋅⋅   ⋅⋅⋅   ⋅⋅⋅   ⋅⋅⋅   ⋅⋅⋅   ⋅⋅⋅ 
B2-A10 4.22(2) 0.71(6) 1.6(3) 14.3(1.2) 10.3(1) 0.29(3) 35(7) 5.0(9) 2.1(4) 4.8(2)
B3-A1 3.15(2) 0.08(8) 0.7(*) 13.9(*) 13.9(1) 0.0(*)  ⋅⋅⋅  2.2(6)  ⋅⋅⋅  12.4(*)
C-A1 4.08(1) 0.16(1) 9.3(6) 10.4(0.7) 8.0(2) 0.0(*) 84(17) 7.4(6) 1.3(3) 4.2(2)
C-A2 3.97(2) 0.36(4) 2.7(6) 12.0(1.2) 8.8(2) 0.13(2) 15(3) 4(1) 4.5(9) 4.1(2)
C-A3 3.98(1) 0.29(2) 9.3(5) 10.7(0.8) 8.0(2) 0.10(1) 68(14) 13(1) 2.8(6) 3.9(2)
F-A1 4.82(4) 0.8(1) 1.5(7) 15.3(2.0) 7.4(2) 0.32(5) 31(6) 4(2) 1.7(3) 1.6(1)
F-A2 4.31(2) 0.58(5) 1.9(4) 15.6(1.6) 9.9(2) 0.23(2) 35(7) 4(1) 1.8(4) 3.4(1)
F-A3 4.31(2) 0.43(3) 2.6(4) 16.8(1.8) 9.8(2) 0.16(2) 24(5) 4.1(8) 2.5(5) 3.0(1)

Notes. Uncertainties are given in brackets beside values and show the uncertainty in the last digit of the value, with the exception of the TK uncertainties which are given in K units. A (*) indicates that the uncertainty in the value at this position were large, even though in most cases the values themselves are reasonable. A good HFS fit was not found at the peak position of NH3 clump B2-A9.

Download table as:  ASCIITypeset image

In Oph B (see Figure 7(a)), the vLSR of NH3 emission has little internal variation, with a mean vLSR = 3.96 km s−1 and an rms of only 0.24 km s−1. An overall gradient is seen across Oph B1 and B2, with smaller vLSR values (3.2 km s−1) at the southwest edge of B1 increasing to 4.6 km s−1 at the most eastern part of B2. Correspondingly, B1 has a characteristic velocity somewhat less than the average (3.8kms−1 ± 0.1 km s−1), while the B2 vLSR is slightly greater (4.1 ±  0.2 km s−1). The filament connecting B1 and B2 is kinematically more similar to B2, but there is no visible discontinuity in line-of-sight velocity of the lines. B3 has the lowest vLSR in Oph B, with an average velocity of 3.3kms−1 ± 0.2 km s−1, or 0.7 km s−1 less than the average of the group. This large difference, greater than that between the mean vLSR values of Oph B and F, suggests that B3 may not be at the same physical distance as the rest of Oph B. The change in velocity between B2 and B3 occurs over a small projected distance (∼30'', or ∼3600 AU). There is some indication that B2 and B3 may overlap along the line of sight, as the determined vLSR values in B2 immediately south of B3 are less than those to the east and west, as might be expected if lower velocity emission is also contributing to the line at that location (as discussed further below, the Δv line widths in this area are larger than the average, as would be expected if unresolved emission from two different velocities is contributing to the observed line).

Oph C (see Figure 8(a)) has a mean vLSR = 4.01 km s−1 with an rms of only 0.07 km s−1. A small velocity gradient of ∼0.4 km s−1 is evident in Oph C, with a minimum line-of-sight velocity of ∼3.7 km s−1 in the southeast, increasing to a maximum of ∼4.1 km s−1 in the northwest. A similar gradient was noted in N2H+(1–0) observations by André et al. (2007), and may be indicative of rotation. The identification of two NH3 clumps in the region, however, could also be indicative of two objects with slightly different vLSR. A slight decrease in vLSR values is seen in the NH3 extension to the north, with C-N2 associated with emission at a slightly lower vLSR than the mean.

Oph F (see Figure 9(a)) has a mean vLSR = 4.34 km s−1 and an rms of only 0.26 km s−1 over the region containing the bulk of the NH3 (1,1) emission and two of the three identified NH3 peaks, despite the presence of four protostars. The filament extending toward the northwestern NH3 peak gradually increases in vLSR, but only by ∼0.2 km s−1. With higher sensitivity, single-dish data show that outside this area vLSR drops to ∼3.7 km s−1. There is clear evidence for two velocity components along the line of sight at the peak position of F-A1, with a secondary component at ∼3.7 km s−1. André et al. (2007) also find two velocity components near F-A1 in N2H+observations. For the brighter component, their N2H+ data agree with our NH3 data, but for the secondary component they find a higher vLSR = 4.1 km s−1. Some blue asymmetry is found in the line profiles of F-A2 and F-A3, but this is more likely due to the complicated velocity structure of the core rather than infall motions (comparison with an optically thin tracer at this position is necessary to confirm infall).

In summary, vLSR varies little across any of the Cores (rms <0.24 km s−1), and additionally varies little between them (maximum mean difference ∼0.38 km s−1). Some small gradients were found in the larger Cores (i.e., on scales larger than the individual NH3 clumps) which may be indicative of rotation.

Figures 7(b), 8(b), and 9(b) show the line Δv for Oph B, C, and F obtained from the combined, smoothed and regridded data. As stated above, the line widths have been corrected for the resolution of the spectrometer (0.3 km s−1). In the few cases where the returned FWHM from the fits is similar or equal to the resolution, we set the corrected FWHM to the thermal line width (see Appendix A for further discussion).

Line widths range from Δv ≲ 0.1 km s−1 to ∼1.2 km s−1 in both Oph B and F, with a minimum line width of 0.08 km s−1 in B3 and a maximum of 1.37 km s−1 in B2. Line widths in Oph C range from 0.11 km s−1 to 0.70 km s−1. The rms variations Δv and vLSR in Oph B and C are similar (0.2 km s−1 and 0.1 km s−1, respectively), while in Oph F the rms in Δv is much larger than the rms in vLSR (i.e., 0.3 km s−1 compared to 0.12 km s−1, respectively).

The extended emission in Oph B is dominated by highly nonthermal motions (mean Δv = 0.83 km s−1), shown in Figure 7(b). Several localized pockets of narrow line width are found embedded within the more turbulent gas. The single Oph B3 clump (B3-A1) and B2-A7 are both characterized by extremely narrow Δv (i.e., 0.08 km s−1 and 0.33 km s−1 respectively). As mentioned above, the line emission broadens from B3 to B2 over only ∼30'' (3600 AU) to Δv ∼ 1.4 km s−1, possibly due to line blending along the line of sight if B2 and B3 overlap in projection at these locations. We find small Δv ∼ 0.5 km s−1 toward the southeastern edge of the mapped region in B1 near NH3 clumps B1-A3 and B1-A4, but the line width minimum does not coincide with either core. Another region of low Δv ∼ 0.6 km s−1 is coincident with the NH3 clumps B2-A1 and B2-A2. A final Δv minimum, Δv ∼ 0.4 km s−1 is found between the two eastern protostars in B2. Note that the protostars in Oph B are all associated with smaller Δv than the average value for the Core, but none are coincident with a clear local minimum in line width.

The maximum line width in Oph C, Δv = 0.70 km s−1, is half that found in Oph B. Figure 8(b) shows that most of the emission in C is narrow, with a minimum Δv = 0.11 km s−1, similar to the extremely narrow lines found in B3. These narrowest line widths are found centered on NH3 clump C-A1 in a band perpendicular to the elongated direction of Oph C and are coincident with the highest velocity emission, and are thus not coincident with the NH3 integrated intensity maximum nor the continuum emission peak. Curiously, if C-A1 and C-A3 are indeed physically distinct clumps, we would expect the broadest lines between them due to overlap, but instead find narrow lines at this location. The largest line widths are found at the edges of the integrated intensity contours.

Oph F is characterized by moderately wide line emission more similar to that found in Oph B, with a mean Δv = 0.63 km s−1 (see Figure 9(b)). Line widths associated with the central NH3 clumps F-A2 and F-A3 are smaller than the mean (i.e., 0.6 km s−1 and 0.3 km s−1 respectively), while clear minima in line width (Δv ∼ 0.33 km s−1) are found at the locations of the two central protostars. Line widths along the Core extending to the northwest integrated intensity peak broaden to larger values (∼1.2 km s−1), but at the tip NH3 clump F-A1 is associated with Δv = 0.4 km s−1 in a single 15'' pixel.

Overall, the observed NH3 Δv in the Cores are generally large, excepting Oph C. We also find regions of localized narrow line emission, some of which are associated with NH3 clumps. Line widths near protostars tend to be smaller than the mean values, but only in Oph F are the protostars coincident with clear minima in Δv.

4.3. Kinetic Temperatures and Nonthermal Line Widths

Following Mangum et al. (1992), among others, we use the returned τ, Δv and line brightnesses of the NH3 (1,1) and (2,2) lines in each pixel to calculate the kinetic temperature of the gas. Details of our calculations can be found in Appendix A. Propagating uncertainties from our hyperfine structure fitting routine gives typical uncertainties in TK of ∼1 K. The mean, rms, minimum, and maximum kinetic temperatures are given for Oph B, C, and F in Table 7. Figures 7(c), 8(c), and 9(c) show the kinetic temperatures calculated across the Oph Cores.

In Oph B, we find a mean TK = 15.1 K with an rms variation across the entire Core of only 1.8 K. Most NH3 emission peaks are associated with lower than average gas temperatures, but only a few are coincident with clear TK minima. The lowest temperatures in the Core, TK = 13.7 K and 12.4 K, are found in southern B1 toward B1-A3 and B1-A4, respectively. These low temperatures are found at the same location as the detection of single dish C2S emission. Gas temperatures appear colder (TK ∼ 13 K) toward the center of Oph B2, but the minimum temperature, TK = 12.4 K, is not coincident with an NH3 clump. Instead, the lowest temperatures are found directly between the NH3 clumps B2-A4, B2-A5, and B2-A6, and closer to the central submillimeter clump.

Oph C is the coldest of the observed Cores, with a mean TK = 12.8 K and a similar rms variation (1.6 K) as in Oph B. The central region is effectively at a single low temperature TK = 10.6 K, with the lowest values found near the emission peaks C-A1 and C-A3 in the northwest and southeast, while the gas temperature of the northern clump, C-A2, is consistent with the average.

Oph F is characterized by the highest temperatures of the observed Cores, with a mean TK = 16.6 K, slightly warmer than Oph B, and with a large rms variation of 3.2 K. No clear minima in gas temperature are observed near any of the NH3 clumps, protostars or continuum peaks identified in the region. The protostar associated with NH3 clump F-A2 is coincident with a temperature maximum in a single 15'' pixel.

The gas temperatures traced by NH3 emission in the Oph B and F Cores are consistently higher than those found in isolated dense clumps. For example, all five of the starless clumps surveyed by Tafalla et al. (2002) were found to have a constant gas temperature TK = 10 K determined through an analysis similar to that done here. Two recent studies of NH3 emission in dense clumps in the Perseus molecular cloud and the less active Pipe Nebula at 32'' resolution also found slightly lower temperatures than those found here, with a median TK = 11 K in Perseus (Rosolowsky et al. 2008a) and a mean TK = 13 ± 3 K for ≲1 M clumps in the Pipe Nebula (Rathborne et al. 2008, we note that one object in the Pipe is warmer than the typical TK found in Oph B and F). In a sample of NH3 observations toward Galactic high mass star forming regions, Wu et al. (2006) found a mean TK = 19 K. The mean kinetic temperatures found in Oph B and F are also slightly greater than the median TK = 14.7 K found in a survey of NH3 observations by Jijina et al. (1999), although their analysis showed that the median temperature of dense gas in clusters was significantly higher, TK = 20.5 K, than in nonclustered environments where the median TK = 12.4 K. Since L1688 is a clustered star forming environment, it is not unreasonable to expect temperatures higher than those in more isolated regions, given the Jijina et al. results.

Evidence for an extremely cold temperature of 6 K (obtained through observations of H2D+(110–111), which likely probes denser gas than NH3 (1,1) and (2,2)) was recently found for the nearby Oph D core (Harju et al. 2008). In addition, while the mean gas temperature TK = 12 K in Oph C, temperatures in the highest column density gas drop to 10 K, similar to temperatures found in the studies of isolated clumps described above.

We find little difference in average gas temperatures between NH3 clumps and NH3 emission associated with submillimeter continuum emission peaks, and a mean TK increase of only ∼1 K, i.e., similar to our uncertainty in TK, in gas temperatures near protostars (but note only five protostars are associated with emission with sufficient S/N to fit the HFS). The Jijina et al. (1999) survey found that NH3 clumps not associated with identified IRAS sources (presumably protostellar objects) were slightly colder than those with coincident IRAS detections (12.4 K compared with 15.0 K), but the effect was much smaller than temperature differences seen due to association with a cluster.

Given the determined gas temperature TK, we calculate the expected one-dimensional thermal velocity dispersion σT of the gas across the Cores:

Equation (3)

Here, kB is the Boltzmann constant, $\mu _{{{\rm NH_3}}} = 17.03$ is the molecular weight of NH3 in atomic units, and mH is the mass of the hydrogen atom. Similarly, the thermal sound speed cs of the gas can be calculated using a mean molecular weight μ = 2.33.

The nonthermal velocity dispersion σNT is given by

Equation (4)

where $\sigma _{\rm obs} = \Delta v / (2 \sqrt{2 \ln {2}})$. The mean, rms, minimum, and maximum values for both σNT and the nonthermal to thermal velocity dispersion ratio of the gas, given by σNT/cs, are given for each of Oph B, C, and F in Table 7.

Figures 7(d), 8(d), and 9(d) show the resulting nonthermal to thermal velocity dispersion ratio over the Cores. The mean σNT/cs values show supersonic velocities are present. At the limits of the velocity resolution of our data, the smallest observed line widths are consistent with motions being purely thermal in nature.

We find that the mean σNT = 0.35 km s−1 and σNT/cs = 1.5 in Oph B. Across the Core, we find a moderate σNT/cs rms of 0.4. The majority of the gas traced by NH3 in Oph B is thus dominated by nonthermal, mildly supersonic motions. Several NH3 clumps are associated with smaller, but still transsonic, nonthermal motions. Thermal motions dominate the observed line widths in only two well defined locations in Oph B which additionally coincide with NH3 clumps: B2-A7, with σNT/cs = 0.5, and B3-A1, where the observed line width is consistent with purely thermal motions. Otherwise, little difference is seen between the nonthermal line widths of individual clumps and the surrounding gas, with a mean σNT/cs = 1.4 km s−1 for the NH3 clumps.

In contrast, Oph C has a mean σNT = 0.14 km s−1. Consequently, the mean σNT/cs = 0.6 with an rms of only 0.2, showing that thermal motions dominate the observed line widths over much of the Core. The minimum nonthermal line width is found associated with C-A1, which is consistent with purely thermal motions within our velocity resolution limits. On average, nonthermal motions in Oph F are also similar to the expected thermal values, with a mean σNT/cs = 0.9, but with a larger spread around the mean (0.6 rms) and a maximum value (σNT/cs = 2.2) similar to that found in Oph B (σNT/cs = 2.5). The lowest values are found toward F-A2 and F-A3 and the nearby protostars.

The nonthermal NH3 line widths we measure are similar to those recently found for N2H+(1-0) emission in the Cores at larger angular resolution (∼26''), where the mean σNT/cs = 1.6 ± 0.3 in Oph B, 0.9 ± 0.2 in Oph C, 1.5 ± 0.8 in Oph F (André et al. 2007), and are less than those found in DCO+ emission (σNT/cs ∼ 2 in B2, ∼1–1.5 in B1, B3, C, and F (Loren et al. 1990).

4.4. Column Density and Fractional Abundance

4.4.1. NH3

Given Δv, τ, and Tex from the NH3 (1,1) line fitting results, we calculate the column density of the upper level of the NH3 (1,1) inversion transition. We then calculate the NH3 partition function to determine the total column density of NH3 following a revised calculation based on Rosolowsky et al. (2008a). Relevant equations are given in Appendix A.

We also calculate the H2 column density, N(H2), per pixel in the cores from 850 μm continuum data using

Equation (5)

where Sν is the 850 μm flux density, Ωm is the main-beam solid angle, μ = 2.33 is the mean molecular weight, $m_{\rm{{H}}}$ is the mass of hydrogen, κν is the dust opacity per unit mass at 850 μm, and Bν(Td) is the Planck function at the dust temperature, Td. We take κν = 0.018 cm2 g−1, following Shirley et al. (2000), using the dust model from Ossenkopf & Henning (1994) which describes grains that have coagulated for 105 years at a density of 106 cm−3 with accreted ice mantles and incorporating a gas-to-dust mass ratio of 100. The 15'' resolution continuum data were regridded to 15'' pixels to match the combined NH3 observations. The dust temperature Td per pixel was assumed to be equal to the gas temperature Tk derived from HFS line fitting of the combined NH3 observations. This assumption is expected to be good at the densities probed by our NH3 data (n ≳ 104cm−3), when thermal coupling between the gas and dust by collisions is expected to begin (Goldsmith & Langer 1978). If these temperatures are systematically high, however, then the derived N(H2) values are systematically low. There is a ∼20% uncertainty in the continuum flux values, and estimates of κν can additionally vary by ∼3 (Shirley et al. 2000). Our derived H2 column densities consequently have uncertainties of factors of a few.

Due to the chopping technique used in the submillimeter observations to remove the bright submillimeter sky, any large-scale cloud emission is necessarily removed. As a result, the image reconstruction technique produces negative features around strong emission sources, such as Oph B (Johnstone et al. 2000a). While the flux density measurements of bright sources are likely accurate, emission at the core edges underestimates the true column. We thus limit our analysis to pixels where Sν ⩾ 0.1 Jy beam−1, though the rms noise level of the continuum map is ∼0.03 Jy beam−1. For a dust temperature Td = 15 K, this flux level corresponds to N(H2) ∼ 6 × 1021 cm−2.

Using the calculated H2 and NH3 column densities, we have calculated per pixel the fractional abundance of NH3 relative to H2, X(NH3) = N(NH3)/N(H2) for each Core. The results of these calculations are shown in Figures 78, and 9, which show the H2 column density derived from submillimeter continuum data, N(H2) and the fractional NH3 abundance, X(NH3). The mean, rms, minimum, and maximum of the derived column density and fractional abundance in each Core are given in Table 7, while specific values for identified NH3 clumps are listed in Table 8. The N(H2) and consequently the X(NH3) uncertainties given in Table 8 include the ∼20% uncertainty in the submillimeter continuum flux values only; uncertainty in κν is not taken into account.

Oph B has a mean NH3 column density of 4.6 × 1014 cm−2, with the highest N(NH3) values (maximum N(NH3) = 12 × 1014 cm−2) found in B1. Two peaks in NH3 column density are found in B1 which correspond closely with the integrated intensity maxima, but are offset from the B1 NH3 clumps. In B2, the column density also generally follows the integrated intensity contours, with lower values overall than in B1. The highest NH3 column in B2 of N(NH3) = 8.5 × 1014 cm−2 is found toward B2-A5. The highest opacity in B2, τ = 4.7, was found associated with B2-A7, but the NH3 column density at this location is similar to the core average. Small column density increases are seen at other NH3 intensity peak locations. The NH3 extension connecting B1 and B2 is characterized by similar NH3 column densities to those at the edges of the core with no obvious N(NH3) maxima.

The discrepancy between the bright NH3 and faint submillimeter continuum emission in B1 indicates a high NH3 fractional abundance relative to B2, with fractional abundances a factor of ≳2–3 higher than the typical values of X(NH3) ∼ 10−8 within B2. This is shown in Figure 7(f). Abundance minima are seen in B2, most notably toward B2-A6 and nearby protostars. Despite the higher NH3 column densities in B1 and B2, the lack of submillimeter emission in the NH3 extension connecting the two regions suggests this connecting material has a higher fractional NH3 abundance. Prominent negative features in the continuum data in this extension preclude a quantitative X(NH3) estimate.

The mean and maximum NH3 column densities in Oph C are similar to those found in Oph B (〈N(NH3)〉 = 5.3 × 1014 cm−2 and N(NH3) = 13 × 1014 cm−2, respectively). Oph C contains two N(NH3) maxima. One is coincident with C-A3 and the second is offset to the west by ∼15'' from C-A1. C-A3 is correspondingly associated with a maximum in NH3 fractional abundance (X(NH3) = 2.8 × 10−8), but C-A1 is coincident with an elongated minimum X(NH3) ∼ 1 × 10−9 that extends along the same axis perpendicular to the long axis of the core where the smallest line widths were found. The highest NH3 abundances X(NH3) ∼ 3 × 10−8, are found in the northern extension. The mean abundance in Oph C, X(NH3) = 1.7 × 10−8, is slightly more than half the Oph B average.

In Oph F, the mean N(NH3) = 2.6 × 1014 cm−2 is less than that found in Oph B and C by a factor of ∼2. The maximum N(NH3) = 6.6 × 1014 cm−2 is also significantly less than the maxima in either B or C, and is found in the emission extending to the northwest from the two central NH3 clumps and protostars. The fractional NH3 abundances are also low compared with B and C, with a mean X(NH3) = 1.8 × 10−8 and a maximum X(NH3) = 3.1 × 10−8 found near but not coincident with F-A3.

Studies of isolated starless clumps have determined a wide range of NH3 abundance values for these objects. While in some cases different methods have been used to determine H2 column density values than that performed here, typical observed fractional abundance values in cold, dense regions tend to be on the order of a few ×10−9 to a few ×10−8 (Tafalla et al. 2006; Crapsi et al. 2007; Ohishi et al. 1992; Larsson et al. 2003; Hotzel et al. 2001). Abundances as low as X(NH3) = 7 × 10−10 and 8.5 × 10−10 have been proposed for B68 (Di Francesco et al. 2002) and Oph A (Liseau et al. 2003), respectively. The values found here agree well with previous studies. The wide variations of X(NH3) in the same general environment suggests dramatic differences in the chemical states of the Cores in L1688 (see Section 5).

4.4.2. C2S and HC5N

Similarly, we can calculate the abundance of C2S and HC5N from respective emission detected in the single-dish data, where

Equation (6)

is the column density of the upper state of the observed transition (Rosolowsky et al. 2008a). The values for Aul, g1, and g2 were taken from Pickett et al. (1998) for each transition. Assuming the transitions are optically thin, the observed temperature of the line TMB = [J(Tex) − J(Tbg)]τul. The partition function $Z = \sum _i g_i \exp \left(\frac{-E_i}{kT_K}\right)$ was then used to calculate the total column density of each species as for NH3, with gi and Ei values taken from Pickett et al. (1998). The column densities thus derived are given in Table 5. The molecular column densities derived for C2S in B1 and C (N(C2S) ∼ 1012–1013 cm−2) are similar to results in young starless cores (Tafalla et al. 2006; Rosolowsky et al. 2008a; Lai et al. 2003). The N(HC5N) results agree with previous measurements in the Taurus molecular cloud (Codella et al. 1997; Benson & Myers 1983) and the Pipe Nebula (Rathborne et al. 2008). We calculate molecular abundances as above and find X(C2S) = 3.1 × 10−10 and 1.5 × 10−10 at the C2S emission peaks in Oph B1 and C, respectively. We further find an abundance X(HC5N) = 4.6 × 10−11 at the HC5N emission peak in Oph C.

4.5. H2 Density

Given the determined excitation and kinetic temperatures, Tex and TK, and assuming the metastable states can be approximated as a two level system, we have calculated the gas density n(H2) from the NH3 (1,1) transition following Ho & Townes (1983). We list the mean, rms variation, and range of densities found for each Core in Table 7. Note that this density is effectively a mean density along the line of sight. In general, we find n(H2)∼ a few ×104 cm−3 in all three Cores, with only moderate variation and no clear spatial correspondence with NH3 or continuum clumps. The largest n(H2) values (n(H2) ∼ 8 × 105 cm−3) were found in Oph B2 toward the central continuum clump B2-MM8 (labeled in Figure 2(b)). While these values agree with n(H2) estimates based on NH3 emission in other regions, they are an order of magnitude lower than estimates of Ophiuchus clump densities derived from dust continuum emission studies at similar spatial resolutions (Motte et al. 1998; Johnstone et al. 2000b).

5. DISCUSSION

5.1. Discussion of Small-scale Features

5.1.1. Correlation Between NH3 clumps, NH3 Integrated Intensity, and Dust Clumps

In Section 3, we used clumpfind to identify objects in NH3 emission within the Oph Cores in position and velocity space. The returned NH3 clump locations are generally found at locations of peak integrated NH3 intensity, with the exception of Oph C, in which we found two distinct NH3 clumps offset from the NH3 integrated intensity maximum.

In Oph B, we find poor correlation between maxima of NH3 integrated intensity and thermal dust continuum emission. Since most NH3 clumps are located at integrated intensity maxima, we find NH3 clumps identified through clumpfind do not correlate well with dust clumps. Continuum dust emission is a commonly used surrogate tracer of gas column density. The observed flux is a function of the dust emissivity (κν) and temperature (Td). If the dust in B1, for example, was colder than that in B2, the same column of dust would produce less emission. In Section 4.3, we calculated H2 column densities assuming the dust and gas are thermally coupled. For Oph B, we found the H2 column density closely followed the observed continuum emission under this assumption (see Figure 7(a) versus Figure 2(b)). The dust and gas, however, may not have the same temperature. If the dust is colder than the gas by a small amount (Td = 10 K compared with TK = 15 K, for example), the true column density of H2 could be larger by a factor of ∼2 along that line of sight. If these temperature differences occur on small enough scales, e.g., at the NH3 clump positions, they could explain the discrepancy between the locations of dust clumps and NH3 clumps. Thermal coupling of gas and dust is most likely to occur, however, in the coldest and densest clumps, i.e., exactly where we do not find correspondence between the dust and gas tracers.

A more likely cause of the offset between dust and NH3 emission is fractional abundance variation of NH3 in the Oph Cores. If the column densities determined from the dust emission are accurate, then most dust clumps are associated with X(NH3) minima. Within B1 and B2, we find variations in X(NH3) of ≳2 on length scales similar to the NH3 clump sizes. Models of nitrogen chemistry in dense regions suggest that a long timescale, greater than the free-fall time, is required for molecules such as NH3 to achieve steady state values, but that during gravitational collapse X(NH3) begins to decrease at densities n ≳ 106 cm−3 (Aikawa et al. 2005; Flower et al. 2006).

The C2S molecule is easily depleted in cold, dense environments with an estimated lifetime of a few ×104 yr (de Gregorio-Monsalvo et al. 2006). It is thus a good tracer of young, undepleted cores (Suzuki et al. 1992; Lai & Crutcher 2000; Tafalla et al. 2004). The detection of C2S in southern Oph B1 and Oph C therefore suggests that these specific locations are chemically, and hence dynamically, younger compared with other regions. The C2S emission detected in both B1 and C is coincident with or only slightly offset from the integrated NH3 intensity peak, suggesting significant depletion has not yet occurred at those particular locations. This conclusion is further bolstered by the fact that we find higher gas densities (see Section 4.5) in B2 than in B1 or C, and both B2 and F are associated with embedded protostars and are therefore likely more dynamically evolved, i.e., denser. The higher levels of nonthermal motions found in Oph B are at odds with what is expected for an evolved, star forming core, however, and we discuss this further below.

5.1.2. Comparison of NH3 Clumps, Submillimeter Clumps and Protostars

We next compare the mean properties of the dense gas associated with the locations of NH3 clumps, submillimeter clumps, and protostars. We note that given the poor correlation between the NH3 clumps and submillimeter clumps and protostars, the derived physical properties (e.g., TK and σNT) at the submillimeter clump and protostellar locations may be associated with larger-scale gas emission along the line of sight rather than the dense clump gas.

In general, we find only small differences between the mean properties of the dense gas at the peak locations of the NH3 clumps, submillimeter clumps, and embedded protostars. The mean kinetic temperatures for NH3 clumps and submillimeter clumps are nearly equal (∼14 K), and only ∼1.5 K less than the values associated with embedded protostars. This difference is not significant given that the uncertainties in TK are on the order of 1 K. Conversely, excitation temperatures associated with embedded protostars are ∼1.5 K lower than that of NH3 clumps and submillimeter clumps where Tex ∼ 10 K, with uncertainties in Tex also ∼1 K. The line widths of submillimeter clumps tend to be larger than those associated with NH3 clumps by only ∼25%.

Some differences in mean properties between objects are notable. For example, protostars have associated Δv and σNT/cs a factor of 2 narrower than both submillimeter and NH3 clumps. Also, submillimeter clumps and protostars have lower fractional abundances than seen for NH3 clumps by a factor of ∼2. Note, however, that only five protostars are found with NH3 emission strong enough to analyze, as we described in Section 3. For this reason, our comparison sample is limited to protostars that are still associated with significant amounts of gas, where conditions are likely more similar to those found in submillimeter clumps than for more evolved protostars.

5.1.3. σNT/cs in Individual NH3 Clumps

We next look at the nonthermal line widths in the individual NH3 clumps in all three Cores. In general, Jijina et al. (1999) found that nonthermal NH3 line widths in clustered environments are larger than those found in isolated regions. In Figure 10, we plot ΔvNT versus ΔvT for the NH3 clumps in each Core. We omit NH3 clumps B3-A1 and C-A1 where the corrected nonthermal line width is effectively zero. We also show the best fit lines found by Jijina et al. to the relationship between thermal and nonthermal line widths in clustered and in isolated regions. Most of the Oph B clumps lie above the ΔvNT–ΔvT trend for isolated clumps and near the trend for the clumps in clustered regions. The good agreement is somewhat surprising given that the majority of the objects in the Jijina et al. sample were observed with ∼4–8 times poorer angular resolution, while those observed with high angular resolution are high mass star forming regions ∼3–7.5 kpc distant and thus have very low linear resolution (excepting Orion B, at a distance of 420 pc). Even at the small spatial scales probed by our observations, Oph B is characterized by wide line widths that follow the relationship found for larger objects in clustered environments. In comparison, the NH3 clumps in Oph C lie well below the clustered ΔvNT–ΔvT trend, with ΔvNT < ΔvT, and also below the ΔvNT–ΔvT trend seen for objects not associated with a cluster. Two of the three Oph F clumps also fall significantly below the isolated object trend, while the third is more turbulent.

Figure 10.

Figure 10. Nonthermal vs. thermal line widths (FWHM) for individual NH3 clumps in the Oph B, C, and F Cores. The two dashed lines show relationships found by Jijina et al. (1999) for NH3 clumps both with and without an associated cluster. Clumps in Oph B follow the clustered trend, but clumps in Oph C are nearer the relationship for more isolated clumps. Clumps in Oph F scatter across the plot. The dotted line represents ΔvNT = ΔvT.

Standard image High-resolution image

5.2. Discussion of the Cores

5.2.1. Trends with N(H2)

In Figure 11, we plot the distribution of TK, σNT/cs, N(NH3) and X(NH3) with N(H2) (calculated in Section 4.4 assuming Td = TK) in Oph B, C, and F. We additionally analyze Oph B1 and B2 separately to examine any potential differences between the two. In Figure 11(a), we find that TK values in Oph C are nearly universally lower than those in the other filaments, and show a tendency to decrease with increasing H2 column density. As described previously, the other Cores are warmer but also do not show a significant trend with N(H2). We show in Figure 11(b) that Oph B1 and B2 are both consistent with having a constant, mildly supersonic ratio of nonthermal to thermal line widths over all N(H2). Oph C line widths are generally dominated by thermal motions, and σNT/cs decreases significantly at N(H2) ≳ 4 × 1022 cm−2. It is interesting to note that there are few data points at these column densities in the other cores, and no observed decrease in σNT/cs. Figure 11(c) and (d) show that in all cores N(NH3) tends to increase with N(H2), following the general trend that the NH3 emission follows the continuum emission in the cores. The differences between NH3 and continuum clumps are due to small-scale differences in maxima. In addition, the fractional NH3 abundance, X(NH3), tends to decrease with increasing N(H2), although we note that Oph B1 and F have relatively few data points. Figure 11(c) also illustrates the high N(NH3) values in Oph B1 relative to N(H2) compared with values in Oph B2, C, and F.

Figure 11.

Figure 11. (a) Kinetic temperature (TK) in K vs. N(H2) (derived from the 850 μm submillimeter continuum data) in cm−2 units for Oph B1, B2, C, and F. In all plots, points shown are values in 15'' pixels. Temperatures in Oph C are nearly universally lower than those in B1, B2, and F, and show a tendency to decrease with increasing H2 column density. A significant trend in temperature with N(H2) is not obvious in the other Cores. (b) σNT/cs vs. N(H2) in Oph B1, B2, C, and F. The dashed line indicates σNT/cs = 1. Oph B1 and B2 are consistent with a constant ratio of nonthermal to thermal line widths, where σNT/cs>1, over all H2 column densities. Oph C line widths are generally dominated by thermal motions, and σNT/cs decreases significantly at H2 column densities above log(N(H2)) ∼ 22.6. (c) N(NH3) (derived from the NH3 HFS line fitting results) vs. N(H2) for Oph B1, B2, C, and F. NH3 column densities tend to increase with H2 column densities in all Cores. Oph B1 has significantly higher N(NH3) for its relatively low N(H2) values compared with the other Cores. (d) X(NH3) vs. N(H2) for Oph B1, B2, C, and F. While (c) shows the NH3 column densities tend to increase with increasing N(H2), the fractional NH3 abundances appear to decrease with higher H2 column densities.

Standard image High-resolution image

If NH3 is depleting at high densities, as the variation of X(NH3) with N(H2) suggests, NH3 may not trace well the TK in the densest and likely coldest regions. Hence, our assumption of Td = TK may not be valid toward the highest columns and we may be underestimating Td and N(H2) by factors of ∼2. For example, Stamatellos et al. (2007) predicted that Td in the centers of the Oph cores could be as low as ∼7 K. Furthermore, X(NH3) (∝N(H2), see Figure 11) may be overestimated by similar factors. High-resolution multiwavelength continuum observations and models are needed to obtain independent assessments of Td throughout the Oph cores.

5.2.2. σNT/cs in Oph B

Jijina et al. (1999) compiled numerous observations of NH3 in dense gas and found that while most starless clumps are characterized by largely thermal motions, a large fraction of clumps in clusters have σNTT, and that the identification of an NH3 clump as part of a cluster has a larger impact on the observed line widths than association with a protostar. These findings are in agreement with our results in Oph B. The line widths in Oph B, with a mean Δv = 0.89 km s−1 and σNT/cs = 1.5, are significantly wider and more dominated by nonthermal motions than those found in isolated cores. While B2 is associated with at least two embedded protostars and B1 appears starless, when studied separately (shown in Figure 12), both cores have similar σNT/cs distributions.

Figure 12.

Figure 12. Histogram σNT/cs in Oph B1, B2, C, and F. Values plotted are those in 15'' pixels. Oph C is characterized by significantly lower nonthermal to thermal line width ratios than the other cores. When analyzed separately, Oph B1 and B2 are very similar in σNT/cs.

Standard image High-resolution image

If the nonthermal component is caused by turbulent motions in the gas, then it is interesting to consider the turbulence source. In the following, we consider the source of wide lines in this region as being due to "primordial" (i.e., undamped) turbulence, protostar-driven turbulence, bulk motions, or biased sampling.

Firstly, Oph B may have nonthermal motions throughout all the core that are inherited from the surrounding cloud and that have not yet been damped. Since Oph B is associated with a few embedded protostars, it is likely that parts of the Core, at the very least, have been at high density for over a free-fall time tff (∼3 × 105 years for n(H2) ≈ 104 cm−3, ∼ncr for the NH3 (1,1) and (2,2) transitions). Since the dissipation timescale for turbulence is ∼tff (Mac Low & Klessen 2004), it is unlikely that the nonthermal motions from the parent cloud have been retained, if the embedded protostars are indeed physically connected to the Core. Determining the relative velocities of the YSOs compared with the Core vLSR would help to address this question.

Secondly, the embedded protostars in Oph B may be adding turbulence to the Core through energy input associated with mass loss. One outflow has been found in CO emission in the region associated with one of the two protostars near the peak NH3 integrated intensity in B2 (IRS45/GY273 and IRS47/GY279, labeled in Figure 2), with a blue lobe toward the west and a red lobe toward the southeast (Kamazaki et al. 2003). We do not, however, see localized regions of wide line widths associated with these or any protostars in Oph B. In fact, protostars are associated with Δv minima. Additionally, wide line widths are found in B1, where there are no embedded protostars and no known outflows. This suggests that the large nonthermal motions across the Core are not driven by embedded YSOs.

Thirdly, the wide line widths seen in the NH3 emission may be indicative of global infall in the Oph B Core. Given the mean NH3 line width in Oph B2, we calculate a virial mass Mvir ∼ 8 M, assuming a density distribution which varies as ρ ∝ r−2. Given mass estimates of the Cores by Motte et al. (1998; which are uncertain to factors ∼2) and accounting for the different Ophiuchus distance used by the authors (160 pc compared with our preferred value of 120 pc), we find M/Mvir ∼ 5 in Oph B2. Since the mean NH3 (1,1) opacity τ = 1.6 over Oph B, the average individual hyperfine component is optically thin, and we would not expect to see the asymmetrically blue line profiles found in collapsing cores in optically thick line tracers. André et al. (2007) find these spectroscopic signatures of infall motions in B2, with a clear blue infall profile toward our B2-A10 NH3 clump, and profiles suggestive of infall toward B2-A7 and other continuum objects in its eastern half. No evidence of infall motions in the tracers used were found toward central B2, but the lines they used (CS, H2CO and HCO+) may suffer from depletion at the high densities and low temperatures found at this location, masking any infall signature. Oph B1, however, is also characterized by wide NH3 line widths and has M/Mvir ∼ 1. Additionally, spectroscopic infall signatures were found in Oph C toward our C-A3 NH3 clump by André et al., where we find extremely narrow NH3 line widths (similarly, B2-A7 contains the narrowest lines observed in B2). It does not appear, then, that infall motions are the sole contribution to the wide lines we find in Oph B.

Lastly, rather than acting as a tracer of the densest gas in Oph B, the NH3 emission may be dominated in this high density environment by emission from the more turbulent outer envelope of the Core. Modeling of the expected emission given a constant abundance of NH3 would be necessary to determine accurately whether depletion is a factor in Oph B (note that we see evidence for decreasing abundance of NH3 with N(H2); see Figure 11(d)). Alternatively, observations of species which are excited at higher densities than 104 cm−3, such as N2H+ and deuterated species such as N2D+ and H2D+, may better probe the dense clump gas. Moderate resolution (∼26'') resolution observations of N2H+(1–0) (André et al. 2007) find narrower line emission in small-scale N2H+ condensations (which we call "clumps") in Oph B1 and the filament connecting B1 and B2 (σNT = 0.15 ± 0.04 km s−1, or σNT/cs ∼ 0.65 for TK = 15 K), but nonthermal line widths in Oph B2 condensations remain transsonic (σNT/cs ∼ 1.1 < 2 on average). In an upcoming paper, we present H2D+ 110–111 observations in B2, showing that nonthermal line widths of gas at densities of ∼106 cm−3 (approximately the critical density of H2D+, depending on the collisional cross section used) in the core are also transsonic, σNT ∼ 0.25 km s−1, or σNT/cs ∼ 1.4 at 10 K (or ∼1.1 at 15 K; R. K. Friesen et al. 2009, in preparation). Regardless of whether or not the line widths seen in NH3 emission are tracing the highest density gas, the mean nonthermal motions of gas in B2 are greater than typically found in isolated clumps.

5.2.3. σNT/cs in Oph C

Motions in the Oph C Core differ substantially from those found in Oph B. Oph C is likely starless, like Oph B1. In contrast to B1 (and also Oph B2 and F), Oph C is dominated by nearly thermal motions (〈σNT/cs〉 = 0.6), shown in Figure 12. Oph C thus appears less affected by the clustered environment, and is kinematically more alike isolated clumps. NH3 clumps in the Pipe nebula and Perseus molecular cloud have similarly narrow line emission, for example, with the σNT/cs ∼ 1–2 in the Pipe (Rathborne et al. 2008) and intrinsic line widths σv typically less than 0.2 km s−1 in Perseus (Rosolowsky et al. 2008a).

In addition, we see in Oph C a progressive decrease in the magnitude of nonthermal motions as we look at molecular lines that trace increasingly high densities. Nonthermal line widths σNT = 0.20–0.28 km s−1 were determined from DCO+(1–0) observations (albeit with a large 1farcm5 beam) of Oph C (Loren et al. 1990). At their peak emission locations in C, we find σNT = 0.16 km s−1 and 0.17 km s−1 for C2S and HC5N, respectively, from our GBT-only data. These locations coincide with the bulk of the NH3 emission in C, where we find the smallest σNT values (consistent at our velocity resolution with nearly or purely thermal motions), and where recent N2H+(1–0) observations by André et al. (2007) also find σNT ∼ 0.13 km s−1 (their C-MM3 − C-MM6). This trend indicates that turbulent motions have decreased at higher densities.

5.2.4. Oph B3

Oph B3 is an unusual object within Oph B. It was originally detected in DCO+ emission (Loren et al. 1990). It is not readily apparent in dust continuum maps of Oph B, and consequently is only identified as a separate clump by Stanke et al. (2006). This lack of prominence suggests that interesting potential sites of star formation may be overlooked by purely continuum surveys. In a similar result, Di Francesco et al. (2004) found a region of extremely narrow N2H+(1-0) line width in Oph A associated with extended thermal continuum emission but not correlated with a continuum clump. The relatively large difference in line of sight velocity between B3 and the Oph B mean (e.g., greater than that between Oph B and F) suggests that B3 may not be physically associated with the rest of Oph B.

B3 is characterized in NH3 emission by moderately bright, extremely narrow lines which are nearly thermal in width. The lack of continuum emission implies a low H2 column density, leading to an NH3 fractional abundance lower limit similar to the X(NH3) values found for clumps in B1 and C. Interestingly, no C2S or HC5N emission was detected in B3, suggesting that B3 may be in a later evolutionary state than B1 and C. Alternatively, if B3 is physically distinct from the rest of Oph B, its initial chemistry may have differed.

The kinetic temperature associated with the B3-A1 NH3 clump is 13.9 K. This is lower than the average TK in Oph B but not particularly cold compared with other NH3 clumps in Oph B. The uncertainty in this value is large due to the low signal-to-noise ratio of the NH3 (2,2) line in the combined data, but we find a slightly lower value, TK = 12.0 ± 0.3 K, from the single-dish data alone. This temperature difference is not large enough to account for the lack of submillimeter emission at the B3 peak if the column density of material was similar to that found in B1 or B2.

Based on the flux observed by Stanke et al. (2006) at the location of B3 (their MMS-108) and the gas temperature derived from our observations, we find a total clump mass for Oph B3 of M ∼ 0.4 M, which is within a factor of ∼2 (i.e., within uncertainties) of the virial mass calculated using the observed NH3 line width and core radius from Table 5 (M ∼ 0.2 M). Accordingly, if it collapses, Oph B3 may form a very low mass star or brown dwarf, depending on how much material passes into a compact protostar. Future sensitive large format millimetre array detectors, such as SCUBA-2, should easily detect objects like B3 at greater than 3σ levels of confidence with relatively short integration times.

5.3. Implications for Clustered Star Formation

Comparisons of the mean values of many parameters found in each of the Cores (Tables 6 and 7) show significant differences between Oph B and C. In particular, Oph C is characterized by significantly narrower line widths and lower kinetic temperatures. This distinction is further illustrated in Figures 11 and 12. An object like Oph C (and perhaps Oph B3) resembles isolated clumps in the close correspondence between the NH3 line map and the dust continuum map, and in the evidence that the peak of the intensity map coincides with local minima in both TK and Δv. Such a core would fit well the idea that clustered star formation is just a spatially concentrated version of isolated star formation, with smaller, denser star forming clumps packed closer together than in isolated regions. The Oph B Core shows a remarkably different behavior. Oph B contains a larger number of NH3 clumps than Oph C or F, and there is much less correspondence between the locations of the NH3 and continuum clumps. The NH3 clumps show modest contrast with interclump gas in their intensity, and little contrast in their velocity or their line width. The greater amount of fragmentation of Oph B than found in C or F may be related to the relatively higher levels of turbulence in the Core, which are significantly greater than typically found in isolated regions. Thus, dense NH3 gas in Oph B does not resemble the dense gas in regions of isolated star formation, and this raises the issue whether this presents a different "initial condition" for clustered star formation.

6. SUMMARY

We have presented combined single-dish and interferometer NH3 (1,1) and (2,2) observations of the Oph B, C, and F Cores in the clustered star-forming Ophiuchus molecular cloud. We additionally present single-dish C2S (21–10) and HC5N (9–8) observations of the Cores. Our main results can be summarized as follows.

  • 1.  
    While the large-scale features of submillimeter continuum emission and NH3 (1,1) integrated intensity appear similar, on 15'' scales we find significant discrepancies between the dense gas tracers in Oph B, but good correspondence in Oph C and F. We find poor correspondence in Oph B between continuum clumps and NH3 clumps identified with 3D clumpfind, with only five of fifteen NH3 clumps located within 30'' (3600 AU) of a dust clump. This is in contrast with previous findings of extremely high spatial correlation between the two gas tracers in isolated, low-mass starless clumps.
  • 2.  
    We find that vLSR varies little across any of the Oph Cores, and additionally varies by only ∼1.5 km s−1 between them.
  • 3.  
    Overall, the observed NH3 line widths in the Cores are generally large, and often slightly supersonic. We also find regions of localized narrow line emission (Δv ≲ 0.4 km s−1), some of which are associated with NH3 clumps. The larger line widths in Oph B (〈Δv〉 = 0.8 km s−1) agree with previous findings for clumps in clustered regions. Line widths in Oph C, however, decrease to nearly thermal values which are more representative of typical isolated clumps.
  • 4.  
    The derived kinetic temperatures of the gas are remarkably constant across Oph B. Kinetic gas temperatures in both B and F are warmer (〈TK〉 = 15 K) than generally found in isolated star forming clumps, but are consistent with temperatures determined for clumps in clustered environments. The center of Oph C shows a minimum TK ∼ 9 K, similar to previous results in isolated clumps.
  • 5.  
    We find no significant difference in TK between NH3 clumps, submillimeter clumps (〈TK〉 = 14 K) and protostars (〈TK〉 = 15.5 K). Most other physical parameters have similarly insignificant variations between the objects. We do find that protostars are associated with significantly smaller line widths (〈Δv〉 = 0.4 km s−1) with approximately equal contribution from thermal and nonthermal motions.
  • 6.  
    We have determined NH3 abundance values toward the Cores, and find they agree with previous estimates of X(NH3) in cold, dense environments. Single-dish observations of C2S and HC5N resulted in only a few detections, and derived column densities (∼1012–1013 cm−2) similar to those found in other molecular clouds.
  • 7.  
    It is unlikely that the wide line widths observed in Oph B are due to pervasive turbulent motions inherited from the parent cloud if the already formed embedded protostars are physically connected to the Core. We find no evidence of influence by the protostars on the gas motions, i.e., through local increases in line widths or gas temperatures. The NH3 abundance in Oph B2 appears to decrease with increasing gas density. We therefore suggest that the NH3 emission is biased, i.e., due to depletion at high inner densities, and is therefore not tracing the densest gas in the Core. This may explain the differences between the locations of NH3 and continuum clumps. Consequently, the gas temperatures may be lower in the center of Oph B than given by the NH3 line ratios.

We thank the anonymous referee for thoughtful comments which improved this paper. We thank H. Kirk for providing SCUBA maps of the regions observed and D. Johnstone for useful discussions. We thank the observatory staff at the ATCA, VLA, and GBT for their assistance in making our observations successful. In particular, we thank J. Ott for aid in setting up the ATCA observations and J. Lockman for providing scripts to grid OTF data from the GBT. The National Radio Astronomy Observatory is a facility of the National Science Foundation operated under cooperative agreement by Associated Universities, Inc. The Australia Telescope is funded by the Commonwealth of Australia for operation as a National Facility managed by CSIRO. The James Clerk Maxwell Telescope is operated by the Joint Astronomy Centre on behalf of the Particle Physics and Astronomy Research Council of the United Kingdom, the Netherlands Association for Scientific Research, and the National Research Council of Canada. R.K.F. acknowledges financial support from the University of Victoria and the National Research Council Canada Graduate Student Scholarship Supplement Program. We also acknowledge the support of the National Science and Engineering Research Council of Canada.

Facilities: ATCA, GBT, VLA

APPENDIX A: DETERMINING PHYSICAL PARAMETERS FROM HFS LINE FITTING RESULTS

A.1. Kinetic Temperature

Since the metastable states across K-ladders are coupled only by collisions, and if we neglect the upper (JK) nonmetastable states, then the populations of molecules in the metastable states can be described by the Boltzmann equation.

Equation (A1)

where Trot(J', K'; J, K) is the rotational temperature relating the populations in the (J, K) and (J', K') states, and ΔE(J', K'; J, K) is the energy difference between the two states. For (J', K') = (2, 2) and (J, K) = (1, 1), ΔE(2, 2; 1, 1)/k = −41.5 K. The state statistical weights, g(J, K) and g(J', K'), are equal to 3 and 5 respectively for the NH3 (1,1) and (2,2) states (Ho & Townes 1983). Assuming the molecular cloud is homogenous such that n(J', K')/n(J, K) = N(J', K')/N(J, K), we can solve for Trot(2, 2; 1, 1) using

Equation (A2)

where ΔTA(1, 1, m) and ΔTA(2, 2, m) are the antenna temperatures of the main component of the (1,1) and (2,2) transitions, respectively, τ(1, 1, m) is the opacity of the (1,1) line summed over the main component only, and we assume that the line widths Δν(1, 1) = Δν(2, 2). We then calculate the gas kinetic temperature TK from the rotational temperature following the updated result given by Tafalla et al. (2002) in their Appendix B:

Equation (A3)

A.2. Column Density

We calculate the column density of the (1,1) transition following Equation (13) of Rosolowsky et al. (2008a):

Equation (A4)

where we have revised the equation from that presented by (Rosolowsky et al. 2008a) to include both parity states of the NH3 (1,1) level in the derived NH3 (1,1) column density. This revision results in greater NH3 (1,1) column densities (and consequently greater total NH3 column densities) than those determined through the original equation by a factor which depends to a small degree on the excitation temperature Tex, and is ∼2 for the regions studied. The statistical weights for the upper and lower states of the (1,1) inversion transition, g1 and g2 are equal. The Einstein A coefficient A(1, 1) = 1.68 × 10−7 s−1 (Pickett et al. 1998), and $\int {\tau (\nu)d\nu } = \sqrt{2\pi } \sigma _v / (c \nu _0) \tau _{\rm tot}$. The total NH3 column density N(NH3) can then be determined by calculating the value of the partition function Z of the metastable states:

Equation (A5)

where the total NH3 column is then N(1, 1) × Z/Z(1, 1). The values of the rotational constants B and C are 298117 MHz and 186726 MHz, respectively (Pickett et al. 1998). The function S(J) accounts for the extra statistical weight of the ortho- over para-NH3 states, with Z = 2 for J = 3, 6, 9, ... and Z = 1 for all other J.

APPENDIX B: CONSEQUENCES OF 0.3 kms−1 VELOCITY RESOLUTION

To fit the hyperfine components of the NH3 (1,1) inversion emission and obtain robust measurements of the opacity and excitation temperature of the line, we required an observational bandwidth wide enough to contain the main and at least two satellite components of the line. To meet this requirement, we were able to obtain velocity resolution across the band of only 0.3 km s−1 due to the present correlator capabilities at the VLA. Here, we investigate the effect of the relatively low spectral line resolution on our determined line widths and kinetic temperatures by creating model spectra, including random noise with an rms value equal to that in our observed spectra. The spectra are then convolved by a finite spectral resolution and resampled. We then fit the resulting spectra with our custom HFS routine.

NH3 (1,1) spectra were modeled given an excitation temperature Tex and opacity τ typical of the regions observed, while the line width Δv was varied. The line intensity at each velocity is then given by Equations (1) and (2). The amplitudes of the 21 components of the NH3 (2,2) line were then determined in a similar manner, with the relative intensities of each component given by Kukolich (1967). The relative intensity of the (2,2) line at a given TK was determined using Equation (A2), by first calculating the associated rotational temperature Trot (reversing Equation (A3)).

While the observational resolution does significantly impact some of the returned parameters, we show that the derived kinetic temperatures in this work are robust.

B.1. Line Widths

Figure 13 plots the returned Δv against that of the model spectrum for 0.1 km s−1 and 0.3 km s−1 resolution for a range of line widths. For both resolutions, the returned line width is greater than the actual for small values, with the 0.3 km s−1 resolution showing a significantly greater effect. The trends, however, follow that expected for addition in quadrature of the true line width and the resolution, i.e. $\Delta v_{\rm obs} = \sqrt{\Delta v_{\rm line} + \Delta v_{\rm res}}$ with some additional offset at small values and larger scatter due to poor resolution. For the mean values found in this study, the corrected line widths consequently reflect the true line widths, but at small widths the corrected values still overestimate the true value. For this reason, we cannot accurately state the true widths of lines where returned (uncorrected) values Δv ≲ 0.35 km s−1, but do show that these uncorrected line widths are consistent with purely or nearly thermal values.

Figure 13.

Figure 13. Input vs. recovered line widths determined by creating a model spectrum convolved to 0.1 km s−1 and 0.3 km s−1 velocity resolution and subsequently fitting with the NH3 HFS routine. Solid lines show the expected observed line width $\Delta v_{\rm obs} = \sqrt{\Delta v_{\rm line} + \Delta v_{\rm res}}$ for 0.1 km s−1 and 0.3 km s−1.

Standard image High-resolution image

B.2. Opacity

Figure 14 plots the returned opacity τ against that of the model spectrum for 0.1 km s−1 and 0.3 km s−1 resolution over a range of line widths. The relatively low spectral resolution of our observations has a significant impact on the returned opacity, with our fitting routine significantly underestimating the opacity by greater than 20% for line widths ≲0.7 km s−1. In addition, the scatter in the returned opacities is large at small line widths, and this affects the uncertainty in the calculated NH3 column density, N(NH3), in Section 4.3. Since N(NH3) depends linearly on the opacity (see Equation (A4)), the uncertainty in the opacity dominates the uncertainty in the column density at small line widths. Since the column density also depends linearly on line width, however, and the returned line width at small values is greater than the true value by a similar relative amount, the returned column density is more accurate than the uncertainties suggest.

Figure 14.

Figure 14. Recovered vs. input NH3 (1,1) opacities as a function of line width Δv for 0.1 km s−1 and 0.3 km s−1 velocity resolution. At small line widths, the returned opacities are significantly smaller than the true values.

Standard image High-resolution image

B.3. Kinetic Temperature

The results from the above tests show that at small NH3 line widths, the returned line widths and opacities can vary significantly from the true values. Regardless, the returned kinetic temperatures from our NH3 line fitting are robust, as shown in Figure 15 (for TK = 15 K). Even at small line widths (Δv ≲ 0.3 km s−1), the kinetic temperatures derived from the fits are accurate within uncertainties of the true TK, with relatively little scatter. At the lowest temperatures found in the Oph cores, TK = 10 K, the results are similar with slightly more scatter at the lowest line widths.

Figure 15.

Figure 15. Recovered vs. input kinetic temperatures TK as a function of NH3 (1,1) line width Δv for 0.1 km s−1 and 0.3 km s−1 velocity resolution. Even at small line widths, the returned kinetic temperature closely agrees with the input value, with little scatter.

Standard image High-resolution image

Footnotes

  • In this paper, we call prestellar objects "clumps" instead of "cores" to avoid confusion with the Ophiuchus "Cores" discussed here.

  • GBTIDL is an interactive package for reduction and analysis of spectral line data taken with the GBT.

  • The NRAO Astronomical Image Processing System.

Please wait… references are loading.
10.1088/0004-637X/697/2/1457