Brought to you by:

A publishing partnership

THE INITIAL CONDITIONS OF CLUSTERED STAR FORMATION. III. THE DEUTERIUM FRACTIONATION OF THE OPHIUCHUS B2 CORE

, , , , , , and

Published 2010 July 6 © 2010. The American Astronomical Society. All rights reserved.
, , Citation R. K. Friesen et al 2010 ApJ 718 666 DOI 10.1088/0004-637X/718/2/666

0004-637X/718/2/666

ABSTRACT

We present N2D+ 3–2 (IRAM), and H2D+ 111–110 and N2H+ 4–3 (JCMT) maps of the small cluster-forming Ophiuchus B2 core in the nearby Ophiuchus molecular cloud. In conjunction with previously published N2H+ 1–0 observations, the N2D+ data reveal the deuterium fractionation in the high-density gas across Oph B2. The average deuterium fractionation RD = N(N2D+)/N(N2H+) ∼ 0.03 over Oph B2, with several small scale RD peaks and a maximum RD = 0.1. The mean RD is consistent with previous results in isolated starless and protostellar cores. The column density distributions of both H2D+ and N2D+ show no correlation with total H2 column density. We find, however, an anticorrelation in deuterium fractionation with proximity to the embedded protostars in Oph B2 to distances ≳0.04 pc. Destruction mechanisms for deuterated molecules require gas temperatures greater than those previously determined through NH3 observations of Oph B2 to proceed. We present temperatures calculated for the dense core gas through the equating of non-thermal line widths for molecules (i.e., N2D+ and H2D+) expected to trace the same core regions, but the observed complex line structures in B2 preclude finding a reasonable result in many locations. This method may, however, work well in isolated cores with less complicated velocity structures. Finally, we use RD and the H2D+ column density across Oph B2 to set a lower limit on the ionization fraction across the core, finding a mean xe,lim ≳ few × 10−8. Our results show that care must be taken when using deuterated species as a probe of the physical conditions of dense gas in star-forming regions.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

The Ophiuchus molecular cloud is a source of ongoing low-mass clustered star formation relatively close to the Sun (d ∼ 120 pc; Lombardi et al. 2008). Through low-resolution DCO+ observations, Loren et al. (1990) discovered a number of dense cores (labeled A through F) in the central Ophiuchus region with masses of a few to several tens of solar masses, which represented a range in evolutionary status given the presence or absence of embedded protostars. Large-scale submillimeter and millimeter observations of Ophiuchus identified several new cores and revealed that the cores were further fragmented into multiple dense clumps, each with masses M ≲ 1 M (Motte et al. 1998; Johnstone et al. 2000, 2004; Young et al. 2006; Sadavoy et al. 2010). Analysis of the clump kinematics suggests that most of them are in virial equilibrium and, hence, not transient objects (André et al. 2007). Recent Spitzer Space Telescope studies of infrared emission toward the cores have characterized in detail the starless and protostellar clump population (Jørgensen et al. 2008; Enoch et al. 2009). These objects are thus ideal locations to study the physical characteristics of dense cores in the process of forming small stellar clusters.

To characterize the physical properties of the dense, cluster-forming Ophiuchus cores, we have previously studied Oph B, C, and F through high-resolution NH3 (1,1) and (2,2) observations (Friesen et al. 2009, hereafter Paper I) and Oph B in both single-dish and interferometer N2H+ 1–0 emission (Friesen et al. 2010, hereafter Paper II). The NH3 and N2H+ observations probe gas at densities n ∼ 103–104 cm−3 and n ∼ 105 cm−3, respectively.

In Paper I, we identified (through the three-dimensional clumpfind structure-finding algorithm; Williams et al. 1994) a number of individual clumps in NH3 emission in the cores, which did not correlate well with those identified in continuum emission. We determined that Oph B and F were warmer on average than typically found for isolated cores (mean TK ∼ 15 K), and we found no trend with temperatures and H2 column density calculated from submillimeter continuum emission. In contrast, Oph C is colder, with temperatures decreasing to a minimum of 9 K at the dust continuum emission peak. Line widths in both the Oph B1 and B2 sub-cores were found to be wide and transonic, despite the presence of three embedded protostars in B2 and none in B1. Oph F, with four embedded protostars, has regions of small and large line widths and complicated line structures. Again similar to isolated objects, NH3 line widths in Oph C are extremely narrow and consistent with being due to purely thermal motions. In an interesting result, the fractional abundance of NH3 with respect to H2, X(NH3), appeared to decrease with increasing H2 column density, N(H2), suggesting that NH3 may be depleted toward locations of high column density and thus may not be tracing the densest core gas. Finally, Paper I revealed that cores with characteristics of both isolated and clustered environments can coexist in a clustered star-forming environment.

In Paper II, we again used clumpfind to identify small-scale structure in N2H+ emission from Oph B and found that the N2H+ clumps matched well NH3 clumps from Paper Iand were consequently again offset from continuum objects. In both studies, little difference was found in the gas properties toward embedded protostars compared with the general core gas. Line widths in Oph B2 remain large at the densities traced by N2H+. In B1, however, N2H+ line widths are significantly narrower than found in NH3 emission. Applying the gas temperatures determined in Paper I, we found that motions in Oph B1 were subsonic with a mean ratio of non-thermal line width to sound speed σNT/cs = 0.7 across the core, while in B2 σNT/cs = 1.3. A clear trend of decreasing X(N2H+) with increasing N(H2) was found toward Oph B2, with emission suggestive of a N2H+ hole toward the continuum emission peak found in high resolution N2H+ 1–0 emission observed with the Australia Telescope Compact Array.

The results of NH3 and N2H+ observations toward Oph B2 in Paper I and Paper II indicate that the highest density gas may best be probed by molecular emission lines excited at higher critical densities or by deuterated molecules, which are expected to be abundant in cold cores where significant depletion has occurred. In cold molecular cloud cores, the deuteration process is initiated by the reaction

Equation (1)

which is exothermic in the forward direction with ΔE = 232 K (Millar et al. 1989). At very low temperatures (≲20 K), the forward reaction dominates, making H2D+ a key molecular ion in the enhancement of deuterated species in a molecular core. CO, which can destroy both H2D+ and H+3 via proton-transfer reactions, is additionally expected to deplete from the gas phase through freeze-out onto dust grains. The deuterium fractionation can thus be elevated orders of magnitude above the cosmic [D]/[H] ratio of 10−5 for H+3, as well as for other species which react with its deuterated forms.

The ortho-H2D+ 111–110 transition at 372 GHz was first detected toward the low-mass young stellar object (YSO) NGC 1333 IRAS 4A by Stark et al. (1999). Multiple detections of H2D+ and doubly deuterated H+3, D2H+, have followed in both starless and protostellar cores (Caselli et al. 2003; Stark et al. 2004; Vastel et al. 2004), including a recent survey of H2D+ emission toward 16 objects by Caselli et al. (2008). Walmsley et al. (2004) show that the H2D+ abundance determined for the dark core L1544 by Caselli et al. (2003) is consistent with a "complete depletion" model, where heavy elements such as C, N, and O (and, consequently, molecules containing these elements) are missing from the gas phase at the highest core densities. In this scenario where no heavy elements are left in the gas phase, H2D+ and multiply-deuterated forms of H+3 are the only gas tracers of the physical conditions of the dense core.

Based on our previous work in Paper I and Paper II, Oph B2 is a good target to investigate the distribution of deuterated species in a more complex star-forming core. To this end, we present in this paper N2D+ 3–2, N2H+ 4–3, and H2D+ 111–110 emission maps of Oph B2. We describe the observations in Section 2 and discuss the observed distributions of N2D+ 3–2, H2D+ 111–110, and N2H+ 4–3 in Section 3. We fit the observed spectra with a single Gaussian (for H2D+) or a multiple-component Gaussian (for N2D+ and N2H+ due to their hyperfine line structure) and discuss the line velocity centroids and widths in Section 4, and also present column density and fractional abundance calculations. In Section 5, we look at the non-thermal line widths in Oph B2 as a function of critical density of the observed emission lines and discuss trends in the deuterium fractionation. Finally, we present a lower limit to the electron abundance in B2 and summarize our results in Section 6.

2. OBSERVATIONS

In the following section, we discuss the observations presented in this paper. We list the species, transitions, and rest frequencies observed in Table 1.

Table 1. Observed Species, Transitions, and Rest Frequencies

Species Transition F'1 F' F1 F Frequency
            (GHz)
N2D+ 3–2 4 5 3 4 231.321906a
N2H+ 4–3 5 5 4 4 372.672494b
H2D+ 111–110         372.421385c

Notes. aGerin et al. (2001). bPagani et al. (2009a). cAmano & Hirao (2005).

Download table as:  ASCIITypeset image

2.1. N2D+ at IRAM

An On-The-Fly (OTF) map of the N2D+ 3–2 line at 231.321 GHz in Oph B2 was made at the Institut de Radio Astronomie Millimétrique (IRAM) 30 m Telescope during the 2007 and 2008 winter semesters using the HEterodyne Receiver Array (HERA; Schuster et al. 2004). HERA is a multi-beam spectral line receiver with nine dual polarization pixels arranged in a 3 × 3 array, with pixels separated by 24'' (approximately twice the 231 GHz 11'' beam full width half-power). The maps were performed with the array tracking the sky rotation and rotated by 9fdg5 with respect to the equatorial system, such that the spacing between OTF lines was 4''. Observations were frequency-switched in-band with a throw of 2 × 6.9 MHz. The Versatile Spectrometer Assembly autocorrelator was used as the backend, with a 20 MHz bandwidth and 40 kHz channel spacing, corresponding to 0.05 km s−1 per channel at 231 GHz.

The data were calibrated to T*A units at the telescope using the Multichannel Imaging and Calibration Software for Receiver Arrays. Further calibration, including baselining and folding of the frequency-switched spectra, was performed using the CLASS software package.8 The data were smoothed by two spectral channels along the frequency axis to improve sensitivity, for a final spectral resolution of 80 kHz or 0.1 km s−1. A second-order polynomial baseline was fitted to the non-line channels and subtracted from the individual spectra before folding the data. The forward (Feff) and beam (Beff) efficiencies of the telescope were interpolated from determined values at 210 GHz and 260 GHz, giving Feff = 0.92 and Beff = 0.58 at 231 GHz. The data were then calibrated to main beam temperature via the relation TMB = (Feff / Beff) T*A. For comparison with the N2H+ 1–0 results in Paper II, we further converted the data to units of Jy beam−1 and convolved the map to a final angular resolution of 18'' to match the N2H+ data.

As the observations were performed in a pooled mode, the 4' × 3' map was observed in three 4' × 1' strips to ensure good sensitivity in potentially limited time. Two-thirds of the map was completed in this manner in winter 2007, while the remaining strip was observed in winter 2008. The sensitivity of the map is consequently not entirely uniform, with the rms noise in the most southern 4' × 1' strip greater (0.15 K (TMB)) than in the rest of the map (0.08 K (TMB)). Most emission was found in the central and northern map sections, so this noise increase does not affect our results significantly.

2.2. H2D+ and N2H+ at the JCMT

Simultaneous observations of the H2D+ 110–111 and N2H+ 4–3 emission lines at 372.421 GHz and 372.673 GHz, respectively, were performed at the James Clerk Maxwell Telescope (JCMT) over the 07A-08B semesters using the 16-receptor Heterodyne Array Receiver Program B-band receiver (HARP-B; Smith et al. 2008). HARP-B is a 4 × 4 pixel array, with array spacing of 30'' and a 2' × 2' footprint. The JCMT beam at 372 GHz is ∼13'' FWHM. We used the ACSIS correlator as the backend, configured to have a 400 MHz effective bandwidth (e.g., wide enough to encompass both lines), tuned halfway in frequency between the two lines, and a frequency resolution of 61 kHz or 0.05 km s−1 at 372.4 GHz. Observations of five separate HARP footprints were performed in position-switching mode to create an undersampled map of H2D+ and N2H+ over Oph B2. The footprints were placed such that the final beam spacing on the sky is 21'' with a coverage of ∼4' × 2'. The fifth footprint provided additional coverage toward the location of peak continuum emission in B2 but has lower sensitivity compared with the other footprints. The observations were performed in band 1 weather only (τ225 GHz < 0.05) and generally while Ophiuchus was above 30° in elevation. The mean rms across the H2D+ map is 0.04 K (T*A) and 0.05 K (T*A) across the N2H+ map. The main beam efficiency factors, ηMB, of the HARP receptors at 372 GHz are not currently well known. At 345 GHz, the mean ηMB = 0.60, with an rms variation of ∼5% between detectors (Buckle et al. 2009). We therefore estimate ηMB ∼ 0.6 with an uncertainty of ∼10%.

Data reduction was performed using the Starlink software package. Each night's data were first checked for noisy or malfunctioning pixels, and these pixels were flagged individually for each integration. A linear baseline was then removed from each integration, and the data were combined into a final cube with 15'' pixels. To increase sensitivity, the data were averaged along the velocity axis by two channels, giving a final velocity resolution of 0.1 km s−1.

3. RESULTS

We first discuss the integrated intensity of N2D+ 3–2, H2D+ 111–110, and N2H+ 4–3 over Oph B2 and compare the species distributions with submillimeter continuum emission and the locations of embedded protostars.

In Figure 1(a), we show the integrated N2D+ 3–2 intensity (convolved to 18'' resolution as described in Section 2.1) in Oph B2 against contours of 850 μm continuum emission observed at the JCMT at 15'' FWHM resolution (original map by Johnstone et al. 2000, combined with all other SCUBA archive data by Jørgensen et al. 2008). Also identified are the locations of 850 μm continuum clumps (Jørgensen et al. 2008) and Class I protostars (classifications and locations taken from Enoch et al. 2009, two of which were previously identified as Elias 32 and 33 (E32 and E33) by Elias 1978; also VSSG 17 and 18 by Vrba et al. 1975). The N2D+ emission is confined to the northern edge of B2 and generally follows the continuum contours in this region, but a very close correspondence is only seen toward the northeast tip of B2. Three maxima of integrated intensity are present, of which two are co-located with continuum clumps while the third is seen to the northwest of the continuum peak. The N2D+ emission is weak toward the continuum emission peak and avoids entirely areas near the protostars.

Figure 1.

Figure 1. (a) Integrated N2D+ 3–2 emission toward Oph B2 (gray scale) in TB units. Data have been convolved to a final 18'' beam (FWHM). Black contours show 850 μm continuum emission in increments of 0.1 Jy beam−1. Squares show the locations of continuum clumps identified by Jørgensen et al. (2008). The continuum clump B2-MM8 is identified. In both plots, stars indicate the positions of Class I protostars (with Elias 32 and 33 identified) and the 18'' beam is shown at lower left. (b) Integrated N2H+ 1–0 emission toward Oph B2 observed with the Nobeyama 45 m Telescope at 18'' (FWHM) resolution (gray scale). Black and gray contours begin at 3 K km s−1 (TMB) and increase by 1 K km s−1. Triangles show the peak locations of N2H+ clumps identified with clumpfind, and the continuum clump B2-MM8 is shown by the square. White contours trace N2D+ emission from (a), beginning at 0.45 K km s−1 (TB) and increasing by 0.15 K km s−1.

Standard image High-resolution image

In Figure 1(b), we plot the N2D+ 3–2 integrated intensity contours over a map of N2H+ 1–0 integrated intensity, observed with the Nobeyama 45 m Telescope at 18'' resolution. The N2H+ observations and analysis were presented in Paper II. We also show the locations of N2H+ clumps, identified through clumpfind in Paper II. Overall, N2D+ emission is found within the extent of N2H+ 1–0 emission. In a similar fashion as the continuum emission, however, we find offsets between the locations of peak N2D+ and N2H+ integrated intensity. In particular, the N2D+ integrated intensity maximum is offset from the integrated N2H+ 1–0 maximum in B2 by ∼20'' to the northeast, ∼1 beam width. The N2H+ 1–0 integrated intensity maximum is itself offset to the east from the continuum peak by ∼22'' (Paper II). The northeastern N2D+ integrated intensity peak, which corresponds well with continuum contours, is also offset from the integrated N2H+ intensity peak by ∼20'' to the east.

We show in Figure 2 H2D+ 111–110 spectra across Oph B2 with overlaid 850 μm continuum contours. We find that significant H2D+ emission is present over much of Oph B2, with a larger extent than found in H2D+ observations toward any other cores so far observed (Caselli et al. 2003; Vastel et al. 2006; Pagani et al. 2009a, 2009b). The H2D+ emission again generally follows the continuum emission but avoids the embedded protostars in B2, similar to N2D+. While relatively strong emission is seen toward the continuum peak, the H2D+ integrated intensity maximum (not shown) is also offset the continuum peak by an estimated ∼20''–30'' (due to the map's undersampling, we are unable to exactly determine the location of maximum integrated intensity). In contrast to the N2D+ emission, the H2D+ integrated intensity maximum is found to the northwest rather than the northeast (although N2D+ emission is also found toward the H2D+ peak). The pixel-to-pixel variations seen in Oph B2 may not be significant, however, due to the ∼5% variation determined for the main beam efficiency factors between different HARP receptors (Buckle et al. 2009). We detect some H2D+ emission in northeast B2, where strong N2D+ emission was found, but do not see a significant integrated intensity maximum. (Note, however, that the HARP footprints do not extend past an R.A. of 16:27:35 and are therefore not sensitive to any emission at the tip of the continuum contours in this region.) Overall, while N2D+ emission appears strongest in the north and east of Oph B2, H2D+ emission is the strongest toward B2's western edge.

Figure 2.

Figure 2. H2D+ 111–110 spectra observed with HARP at the JCMT toward Oph B2 with 15'' FWHM. The gray contours show 850 μm continuum emission observed at the same resolution at 0.1 Jy beam−1, 0.3 Jy beam−1, and 0.5 Jy beam−1. Stars represent Class I protostars, with solid lines indicating the location if within a spectrum box. The velocity scale on the subplots runs from −1 km s−1 to 9 km s−1, while the amplitude ranges from −0.25 K to 0.85 K (TMB, assuming ηMB = 0.6).

Standard image High-resolution image

In Figure 3, we show N2H+ 4–3 spectra across Oph B2 with overlaid 850 μm continuum emission contours. We find bright N2H+ 4–3 emission over most of B2. The greatest integrated N2H+ 4–3 intensity (not shown) is found toward E33, while additional strong N2H+ 4–3 emission is found toward E32 and the 850 μm continuum emission peak. A separate integrated intensity peak is found in northeast B2 (where we also find strong N2D+ 3–2 emission). Overall, while the N2H+ 4–3 emission is generally confined to areas where there is also continuum emission, the strongest N2H+ 4–3 emission is found away from the brightest continuum contours. A comparison between Figure 3 and the N2H+ 1–0 integrated intensity in Figure 1(b) (gray scale) reveals a significant discrepancy between the locations of strong N2H+ 1–0 and 4–3 emission in Oph B. We probe the possible causes of the offset, including the excitation and opacity of the lines, further below.

Figure 3.

Figure 3. N2H+ 4–3 spectra observed with HARP at the JCMT toward Oph B2 with 15'' FWHM. The gray contours, stars, and velocity scale on the subplots are as in Figure 2. The amplitude ranges from −0.3 K to 2.2 K (TMB, assuming ηMB = 0.6). Four locations of interesting spectra, labeled A, B, C, and D, are highlighted by dark outlines.

Standard image High-resolution image

4. ANALYSIS

4.1. Line Fitting

The rotational transitions of the linear molecular ions N2H+ and N2D+ contain hyperfine structure, which is dominated by the interactions between the molecular electric field gradient and the electric quadrupole moments of the two nitrogen nuclei (Dore et al. 2004). Rotational transitions at higher J contain greater numbers of hyperfine components at smaller frequency intervals, i.e., the satellite components are found more closely spaced and can overlap significantly. In particular, the N2H+ 4–3 and the N2D+ 3–2 transitions contain 21 and 25 hyperfine components, respectively, with relative line strengths ≳0.001 (Pagani et al. 2009a, 2009b; Gerin et al. 2001).

For N2H+ 4–3, we calculate the relative line strengths $s_{\it J\,F_1\,F \rightarrow J^\prime \,F_1^\prime \,F^\prime }$ from the Einstein A coefficients $A_{\it J\,F_1\,F \rightarrow J^\prime \,F_1^\prime \,F^\prime }$ and frequencies $\nu _{\it J\,F_1\,F \rightarrow J^\prime \,F_1^\prime \,F^\prime }$ given by Pagani et al. (2009a, 2009b), using

Equation (2)

where μ = 3.4 D is the permanent electric dipole moment, J is the upper rotational quantum number, and [F] = 2F + 1 accounts for the degeneracy of the hyperfine states. For the N2H+ 4–3 transition, 14 components at 8 distinct frequencies make up the main centroid emission and lie within 0.4 km s−1 of each other. At low optical depths the line profile does not differ substantially from a Gaussian profile, while the relative intensities of the satellite lines to the main line component remain small even at large opacities (∼20% at τ ∼ 10). Nevertheless, the satellite components of the line are visible toward several pixels in Oph B (identified as locations A, B, and D on Figure 3, and shown in Figure 4), largely near the embedded protostar E33. At some locations, the line structure is confused with a significant blue emission shoulder, which is not due to any hyperfine structure, but which may be associated with an outflow stemming from either E33 or E32 (Kamazaki et al. 2003; G. White et al. 2010, in preparation). We were thus able to fit well the hyperfine structure toward only one location in the map and constrain the excitation temperature, Tex, and opacity, τ, toward two others.

Figure 4.

Figure 4. N2H+ 4–3 spectra toward four locations in Oph B2, labeled A, B, C, and D in Figure 3: 30'' north of E33 (A), toward E33 (B), 21'' northeast of E33 (C), and ∼42'' northeast of E33 (D). Overlaid on the spectra are fits to the full hyperfine structure of the N2H+ 4–3 line (A, B, and D) and a two-component Gaussian fit (C).

Standard image High-resolution image

We fit the hyperfine structure with a custom Gaussian hyperfine structure fitting routine written in IDL, described in detail in Paper I, assuming equal excitation temperatures Tex for each hyperfine component. The line fitting routine does not take into account the significant overlap between hyperfine components in both the N2H+ and N2D+ transitions, but we find the total N2H+ line opacities are still small where the satellite structure is seen (τ ∼ 2–3) and we therefore do not expect line overlap to introduce substantial errors.

At low opacities, we cannot solve for τ and Tex independently. Given a reasonable estimate for Tex, however, we may attempt to fit the line without visible satellite structure. In Paper II, we determined Tex for N2H+ 1–0 emission across Oph B, and to first order can expect the same Tex to describe the excitation of all rotational levels of N2H+. This assumption was found to be accurate in a recent study of N2H+ and N2D+ emission in L1544, where Caselli et al. (2002b) were able to determine separately the excitation temperatures for N2H+ 1–0, 3–2 and N2D+ 2–1 and 3–2 and found that they were consistent within uncertainties. We find, however, that while most of the data can be fit reasonably well using the N2H+ 1–0 Tex values, visual inspection showed that a simple Gaussian fits the data equally well, or significantly better, in many cases, and the line opacities were not well constrained.

Because multiple components of similar line strength overlap to form the main N2H+ 4–3 component, a Gaussian fit will overestimate the line width, while the returned vLSR will be shifted relative to a full hyperfine structure (HFS) fit. The shift in vLSR can be corrected (ΔvLSR = 0.036 km s−1), and at low opacities, the increase in Δv is small (≲10% for τ < 1, up to ∼25% for τ ∼ 2). In the following sections, we thus discuss only the results of the Gaussian fitting of the N2H+ 4–3 emission unless otherwise stated.

No hyperfine structure was seen above the rms noise in N2D+ 3–2 emission across Oph B. We are therefore unable to fit well the N2D+ 3–2 hyperfine structure solving for both τ and Tex. Here, however, equating the N2D+ Tex with that found for N2H+ 1–0 emission resulted in better constraints on the opacity. The hyperfine fit, using the relative component strengths and velocities from Gerin et al. (2001), is thus more informative than a simple Gaussian fit, and we discuss the results from the full HFS fit in the following sections.

To properly calculate the line width, Δv, we subtracted in quadrature the resolution width, Δvres (0.1 km s−1 for all species), from the observed line width, Δvobs, such that $\Delta v = \sqrt{\Delta v_{\rm obs}^2 - \Delta v_{\rm res}^2}$. The uncertainties reported in the returned parameters are those determined by the fitting routine, not taking the calibration uncertainty into account. The calibration uncertainty does not affect the uncertainties returned for vLSR or Δv. The excitation temperature, column densities, and fractional abundances discussed further below, however, are dependent on the amplitude of the line emission and thus affected by the absolute calibration uncertainty.

The H2D+ 111–110 emission line has no hyperfine structure. We consequently fit the emission at each pixel with a single Gaussian (again in idl) to determine the line amplitude, vLSR and Δv across Oph B2.

4.2. Centroid Velocity and Line Widths

Table 2 lists the mean, rms, minimum, and maximum values of vLSR and Δv for N2H+ 4–3, N2D+ 3–2, and H2D+ 111–110, along with the same results from NH3 (1,1) and N2H+ 1–0 presented in Paper I and Paper II for comparison.

Table 2. Line Properties of NH3, N2H+, N2D+, and H2D+ in Oph B2

Species Transition vLSR       Δv       ncr
    (km s−1)       (km s−1)       (cm−3)
    Mean rms Min Max Mean rms Min Max  
NH3 (J, K) = (1, 1) 4.14 0.19 3.85 4.58 0.87 0.16 0.42 1.11 2 × 104 a
N2H+ J = 1–0 4.04 0.20 3.61 4.43 0.71 0.22 0.38 1.42 1.4 × 105 b
H2D+ $J_{K_{-1}K_1}$ = 111–110 4.01 0.24 3.65 4.50 0.74 0.19 0.51 1.36 ∼105 c
N2D+ J = 3–2 4.02 0.19 3.67 4.44 0.53 0.20 0.25 1.36 8 × 105 d
N2H+ J = 4–3e 4.05 0.22 3.39 4.43 0.69 0.18 0.47 1.14 7.7 × 106 b

Notes. The velocity resolution of the NH3 observations was 0.3 km s−1. All other species were observed with 0.1 km s−1 velocity resolution or better. aCaselli et al. (2002a). bSchöier et al. (2005; ncr calculated using Einstein A coefficients and collision rates γul at 10 K from the Leiden Atomic and Molecular Database). cCaselli et al. (2008). dCalculated using Einstein A coefficients and collision rates for N2H+ 3–2 at 10 K (Schöier et al. 2005). Daniel et al. (2007, see their Figure 17) found that despite small differences in Einstein A coefficients and collision rates, critical densities of N2H+ and N2D+ are similar for transitions other than J = 1–0. eValues calculated from a Gaussian fit due to a lack of visible hyperfine structure. The reported line widths may be greater by ∼10%–25% than the true values; see Section 4.1.

Download table as:  ASCIITypeset image

We show in Figure 5 the N2D+ vLSR and Δv over Oph B2. We only plot values where S/N of the peak line intensity was ⩾5. The mean vLSR across B2 is 4.02 km s−1 with an rms variation of 0.19 km s−1. Lower vLSR values are found toward the western and northeastern regions of the core, with the maximum vLSR = 4.44 km s−1 in the east toward the NH3 (1,1) clump B2-A7 (identified in Paper I and labeled on Figure 5; coincident with the N2H+ 1–0 clump B2-N6 from Paper II). The variation in vLSR does not suggest global rotation of B2. The N2D+ line velocities shift between 3.7 km s−1 and 4.4 km s−1 over ∼40'', or ∼4800 AU, to the west of E32. Nearby, the line vLSR shifts from 3.7 km s−1 to 4.2 km s−1 on either side of the western embedded protostar.

Figure 5.

Figure 5. (a) N2D+ 3–2 line velocity or vLSR in Oph B2. Color scale is in km s−1. In both plots, only results from pixels where S/N of the peak line intensity is ⩾5 are shown, indicated by the solid white contour. Contours show 850 μm continuum emission at 15'' (FWHM) resolution observed with the JCMT, in increments of 0.1 Jy beam−1. The 18'' (FWHM) beam is shown at lower left. Stars indicate the positions of Class I protostars. The locations of the NH3 clump B2-A7 and 850 μm continuum clump 16275-2422 are also shown. (b) Fitted Δv in Oph B2, contours as in (a). Color scale is in km s−1.

Standard image High-resolution image

The mean N2D+ 3–2 line width Δv = 0.53 km s−1 with an rms variation of 0.20 km s−1. In general, the widest line widths are found near E32 and the western protostar (Oph-emb5, Enoch et al. 2009), to a maximum Δv = 1.36 km s−1 in western B2 (but note that the largest Δv values are found only ∼45'' to the northwest of some of the narrowest Δv values). The locations of large line widths correspond well with the locations of dramatic vLSR shifts described above. Away from the protostars, only small variations in the relatively narrow line widths are seen, notably toward the northeast and southwest, with Δv ∼ 0.3 km s−1. Two objects are associated with localized narrow Δv. An 850 μm continuum clump (162725-24273, Jørgensen et al. 2008), shown on the image, was identified where the southwestern narrow line widths are found. In the east, B2-A7 is also associated with an NH3, a Δv minimum of 0.36 km s−1. This is similar to the clump line width in N2H+ 1–0 emission at 18'' resolution, but nearly a factor of two larger than found in N2H+ 1–0 interferometer observations of the clump presented in Paper II.

In Figure 6(a) and Figure 6(c), we show the vLSR determined through Gaussian fitting of the H2D+ 111–110 and N2H+ 4–3 lines, respectively. Visual inspection of the vLSR values shows that the H2D+ emission is similar kinematically to the N2D+ 3–2 values in Figure 5(a), while a small shift in vLSR of ∼+0.2 km s−1 is seen in N2H+ 4–3 emission toward Elias 32 and the 850 μm continuum emission peak and co-located 850 μm continuum emission peak (co-located with the continuum clump B2-MM8). The shift is small but significant given the otherwise excellent agreement in vLSR between the different tracers over B2 (see Table 2).

Figure 6.

Figure 6. (a) and (c) H2D+ 111–110 and N2H+ 4–3 line velocity vLSR in Oph B2, respectively. Color scale is in km s−1. We show only results from pixels where S/N of the peak line intensity is ⩾5. Gray contours show 850 μm continuum emission as in Figure 5. Stars indicate positions of embedded Class I protostars. (b) and (d) H2D+ and N2H+ line width Δv in Oph B2, respectively. Color scale is in km s−1.

Standard image High-resolution image

We show in Figure 6(b) and Figure 6(d) the Δv of the H2D+ and N2H+ lines. The H2D+ line widths are large, with a mean Δv = 0.74 km s−1, and do not vary greatly over Oph B2 (Δv rms = 0.19 km s−1). Slightly greater line widths are found in the western half of B2. Gaussian fits to the N2H+ 4–3 data find that the lines are slightly more narrow, with a mean Δv = 0.69 km s−1. As noted in Section 4.1, Gaussian fits likely overestimate the true N2H+ line width by ∼10%–25%. Wider N2H+ lines are also found toward western Oph B2, while narrower lines are seen toward the submillimeter continuum emission peak and in B2's eastern half. N2H+ lines of average width are found toward the embedded protostars but become wider at pixels directly offset from the infrared sources. In particular, directly to the east of E32 and E33, the N2H+ 4–3 line profiles show a significant blue shoulder with emission extending to vLSR ∼2.5 km s−1. Figure 4(c) shows the result of a two-component Gaussian fit to the line and blue shoulder at location C in Figure 3.

4.3. N2H+ Opacity and Excitation Temperature

Comparison of the N2H+ 1–0 and 4–3 line emission (see Figure 1(b) and Figure 3) shows that while in both cases emission is found over much of Oph B2, the integrated intensity distributions of the two transitions are quite different. It is thus unclear whether the excitation conditions for both transitions would be equal.

Toward several pixels, hyperfine structure is visible. These locations have been labeled A, B, and D in Figure 3, and these spectra are again shown in Figure 4. Using the full hyperfine component fit (overlaid on the spectrum), we were able to determine reasonably well the excitation temperature and opacity of the N2H+ 4–3 line, Tex,4–3 and τ4–3 toward location A, with the resulting best-fit total line opacity τ4–3 = 2.5 ± 1 and Tex,4–3 = 8 ± 1 K. Both Tex,4–3 and τ4–3 are approximately consistent with values found in Paper II for N2H+ 1–0 emission at the same location, where Tex,1–0 = 7.3 ± 0.1 K and τ1–0 = 3.1 ± 0.6. The brightest N2H+ 4–3 line emission, however, and additionally the maximum integrated intensity, is located toward E33 (location B in Figure 3, directly south of A). The intensity ratio of the main N2H+ 4–3 component with an apparent red satellite (line strength only 2× the data rms noise level, but at the correct velocity and with a width that matches the main component, see Figure 4(b)) suggests that the N2H+ 4–3 opacity toward E33 remains moderately optically thick, τ4–3 ∼ 3, and the corresponding Tex,4–3 ∼ 8. The expected line profile given these values is overlaid on the spectrum in Figure 4. The integrated N2H+ 1–0 emission decreases significantly toward E33 (see Figure 1(b) and Paper II). Additionally, the emission is confused by multiple velocity components which preclude robust fits to the 1–0 spectra, so we do not have a good estimate of the N2H+ 1–0 Tex at this location. At location D in Figure 3, satellite emission is again seen, but here the hyperfine fits are confused by significant blue shoulder emission. The best-fit results give Tex ∼ 7.5 K–10 K, with τ ∼ 1 − 3. We overlay the expected line profile for Tex = 8.5 K and τ = 2 on spectrum D in Figure 4. Elsewhere, no hyperfine structure is visible in the N2H+ 4–3 line.

For opacity ratios between the N2H+ 4–3 and N2H+ 1–0 lines, τ4–31–0 ∼ 0.3 and similar line widths, the derived N2H+ column density, N(N2H+), calculated using either transition following Caselli et al. (2002c, Equation (A1)), is equal if both Tex,1–0 and Tex,4–3 ∼ 7 K. At lower opacity ratios, the required Tex also decreases. On average, the N2H+ 1–0 and 4–3 line widths agree, however we note that since we cannot fit well the hyperfine structure, the N2H+ 4–3 widths are an overestimate of the true value by ∼10%–25%. This uncertainty propagates linearly into the column density calculation. With this caveat, an excitation temperature Tex ∼ 7 K agrees well with the mean Tex determined from HFS fitting of the N2H+ 1–0 line in Paper II. We have no column density estimate toward E33 from N2H+ 1–0 emission. Near E32 and E33, however, the greater relative N2H+ 4–3 opacity requires either a higher common Tex ∼ 10–11 K to obtain similar N(N2H+) estimates, which does not agree with our N2H+ 1–0 Tex measurement, or the excitation of the N2H+ 4–3 line must be greater than that of the 1–0 line by ∼2–3 K. This moderate increase in excitation temperature of the higher excitation N2H+ line is not ruled out by our data. Detailed modeling of the N2H+ emission near the embedded protostars using radiative transfer codes would help disentangle the excitation and opacity effects.

4.4. Column Density and Fractional Abundance

We next discuss the column densities of N2D+ and H2D+ derived from the observed line emission. The N2H+ column density was calculated in Paper II.

4.4.1. N2D+

We calculate the total N2D+ column density, Ntot(N2D+), from the integrated intensity of the 3–2 transition following Caselli et al. (2002a) for optically thin emission:

Equation (3)

where W is the line integrated intensity, A is the Einstein spontaneous emission coefficient, gu and gl are the statistical weights of the upper and lower states, respectively, and the equivalent Rayleigh–Jeans excitation and background temperatures are given by Jν(Tex) and Jν(Tbg). The rotational energy EJ = J(J + 1)hB for linear molecules, where the rotation quantum number J = 2 for the lower energy state and B = 38554.717 MHz is the rotational constant (Sastry et al. 1981). The partition function Qrot(Tex) = ∑J = 0(2J + 1)exp(−EJ/kTex), and we use the integrated value, Qrot(Tex) = kTex/(hB) + 1/3. We assume that the excitation temperature Tex is the same for all rotational levels, and use the fitted N2H+ 1–0 Tex from Paper II as an estimate of the excitation temperature of the N2D+ line at each pixel (as discussed in Section 4.1). All other parameters remaining equal, at the average Tex of the N2H+ 1–0 line, a decrease in Tex by 2 K would increase the resulting N2D+ column density by a factor ≲2, while an increase in Tex by 2 K would decrease the resulting N2D+ column density by ≲0.5. We show N(N2D+) across Oph B2 in Figure 7(a).

Figure 7.

Figure 7. (a) Total N2D+ column density N(N2D+) (cm−2) in Oph B2 determined using Equation (3) (gray scale). Values shown have been divided by 1011. In both plots, only results from pixels where S/N of the peak line intensity is ⩾5 are shown, indicated by the thick black contour. Gray contours show 850 μm continuum emission as in Figure 5. The 18'' (FWHM) beam is shown at lower left. Stars indicate the positions of Class I protostars. (b) Fractional abundance of N2D+ relative to N2H+ and RD, in Oph B2. The scale has been truncated to show the variation in central B2; the maximum RD = 0.16 toward the eastern tip.

Standard image High-resolution image

We then calculate the fractional N2D+ abundance per pixel,

Equation (4)

where the H2 column density, N(H2), is calculated from 850 μm continuum emission. The 15'' data were convolved to a final 18'' FWHM resolution to match the convolved N2D+ data. The H2 column density is then given by N(H2) = Sν/[ΩmμmHκνBν(Td)], where Sν is the 850 μm flux density, Ωm is the main-beam solid angle, μ = 2.33 is the mean molecular weight, mH is the mass of hydrogen, κν is the dust opacity per unit mass at 850 μm, where we take κν = 0.018 cm2 g−1, and Bν(Td) is the Planck function at the dust temperature, Td. As in Paper I and Paper II, we expect the gas and dust to be well coupled in Oph B2, and set Td = TK, the gas kinetic temperature determined from NH3 (1,1) and (2,2) inversion line ratios. We refer the reader to the detailed discussion of the uncertainties in the N(H2) values presented in Paper II, but note that the derived H2 column densities have uncertainties of factors of a few. Additionally, due to the negative artifacts introduced into the data from the observational chopping technique (again discussed in detail in Paper II), we limit the 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.

The column density distribution is not similar to the integrated intensity map shown in Figure 1(a). Figure 7(a) shows several small peaks of N(N2D+) in B2 which largely correspond with local minima in the excitation temperature (Tex ∼ 5 K compared with the mean Tex ∼ 7 K, where the uncertainty in Tex is ∼0.4–0.6 K). The N(N2D+) peaks remain within the continuum contours but avoid maxima of continuum emission. We find a mean N(N2D+) = 1.8 × 1011 cm−2 in Oph B2 with an rms variation of 1.3 × 1011 cm−2. The maximum N(N2D+) = 7.1 × 1011 cm−2 is found toward the northeast. If we calculate N(N2D+) assuming a constant Tex = 7 K, the distribution follows the integrated intensity but the mean value and variation remain the same.

The fractional N2D+ abundance (not shown) follows a similar distribution to the N2D+ column density. We find a mean X(N2D+) = 8.2 × 10−12 with an rms variation of 5.1 × 10−12. The maximum X(N2D+) = 2.9 × 10−11 is again found toward northeast Oph B2. Within the continuum emission contours, the lowest measured N2D+ abundances (X(N2D+) ∼ 2 × 10−12) are found toward the continuum emission peak and B2-MM8 clump (emission toward the Elias 33 protostar was not above our S/N limit for the analysis).

4.4.2. H2D+

To calculate the H2D+ column density, N(H2D+), we first estimate the line opacity τ from the H2D+ 111–110 line emission using

Equation (5)

where we must assume a priori an excitation temperature Tex for the transition. An upper limit is given by the gas kinetic temperature TK. For the pixels where the H2D+ S/N >5, we found in Paper I a mean TK = 14.5 K with an rms variation of 1.1 K (slightly less than the mean TK = 15 K found for the entire Oph B core in Paper I), with a minimum TK ≲ 13 K toward the highest N(H2) values. We note, however, that the critical density at which NH3 is excited (ncr ∼ 103–104 cm−3) is lower than that of the H2D+ transition by an order of magnitude, and the NH3 temperatures, or the variation in TK, may not accurately reflect the conditions where H2D+ is excited.

In their recent H2D+ study of several starless and protostellar cores, Caselli et al. (2008) found that the H2D+ Tex was similar to or slightly less than the core TK values, which were estimated through a variety of methods (e.g., continuum observations, NH3 inversion observations as for B2, and multiple transition observations of molecules such as HCO+). Six of the seven protostellar cores in their sample had excitation temperatures which ranged from 9 K to 14 K and similar densities to those calculated for B2 in Paper II. Based on these values, and to better compare our results with Caselli et al., we use a constant excitation temperature Tex = 12 K in Equation (5). Given Tex = 12 K, we find a mean τ = 0.13, with a minimum τ = 0.05 and maximum τ = 0.28, which agrees well with the Caselli et al. results in protostellar cores.

The total column density of ortho-H2D+ is then given by Vastel et al. (2006):

Equation (6)

where gu = 9, Eu/k = 17.8 K, and Aul = 1.08 × 10−4 s−1 for the 111–110 transition (Ramanlal & Tennyson 2004). The integral $\int \tau \mbox{d}v = \frac{1}{2}\sqrt{\pi /(\ln (2)} \tau \Delta v$. We calculate Qrot by reducing the H2D+ level structure to a two-level system following Caselli et al. Since the energy of the first excited state above ground is E/k = 17.8 K and that of the second excited state is E/k = 110 K, we expect this approximation to be valid in Oph B2 given the low temperatures determined in Paper I. The partition function depends on Tex, which we have estimated to be 12 K. In Figure 8, we plot the variation in τ, Qrot, and N(ortho-H2D+) with Tex given a line TMB = 0.5 K and Δv = 0.7 km s−1 (typical values in B2) and have listed the returned values for Tex between 7 K and 15 K in Table 3. While the variation in returned parameters is fairly large with small Tex changes at low Tex values (i.e., factors ∼2 or more around Tex ∼ 7 K), at higher Tex the values vary less. For example, an increased Tex = 14 K rather than Tex = 12 K would result in a decreased N(ortho-H2D+) by ∼20%, while a decreased Tex = 10 K would result in an increased N(ortho-H2D+) by ∼30%. We thus assume an uncertainty of ∼25% in the N(H2D+) values reported here.

Figure 8.

Figure 8. Impact of the assumed excitation temperature, Tex, on returned H2D+ 111–110 line opacity (top), partition function Qrot (middle), and column density N(ortho-H2D+)(bottom), given a line brightness temperature TMB and line width Δv similar to observed mean values in Oph B2. At low Tex, τ and N(ortho-H2D+) vary strongly with Tex, but the effect of changes in Tex at higher temperatures becomes small.

Standard image High-resolution image

Table 3. Impact of Tex on τ, Qrot, and N(ortho-H2D+)

Tex τ Qrot N(ortho-H2D+)
(K)     (1013 cm−2)
7 0.41  9.7 1.60
8 0.27 10.0 1.11
9 0.20 10.2 0.86
10 0.15 10.5 0.71
11 0.12 10.8 0.61
12 0.10 11.0 0.54
13 0.09 11.3 0.48
14 0.08 11.5 0.45
15 0.07 11.7 0.42

Note. Values were calculated for a line strength TMB = 0.5 K and width Δv = 0.7 km s−1.

Download table as:  ASCIITypeset image

To calculate the total H2D+ column density in B2, an estimate of the ortho- to para-H2D+ (o/p-H2D+) ratio is needed. The ground state para-H2D+ transition occurs at a frequency of 1.37 THz; this line is difficult to observe from the ground and may not be excited at the low temperatures found in star-forming cores. Without a direct measurement of the population of the para state, we look to chemical models to predict the o/p-H2D+ ratio, which show that the o/p-H2D+ ratio is directly dependent on o/p-H2. Walmsley et al. (2004) find that the o/p-H2 ratio increases to a steady-state value of ∼5 × 10−4 at T = 10 K and n = 106 cm−3, similar to the density and temperature determined for Oph B2 in Paper I and Paper II. The same models predict o/p-H2D+ ∼0.2–0.3. Using Bonnor–Ebert sphere core models, Sipilä et al. (2010) also predict o/p-H2D+ ratios ≲0.5 at T ∼ 10–15 K in steady state. Their models show that the o/p-H2D+ ratio can dramatically increase to values ∼2–3 at T ≲ 10 K, but these temperatures are unlikely based on our analysis of Oph B2.

It is not clear, however, whether steady-state models accurately represent conditions in star-forming cores like B2. The dynamical timescale of a collapsing core can be much shorter than some of the chemical timescales, leading to large departures of free-fall model predictions from steady-state models (Flower et al. 2006). For example, Flower et al. show that the steady-state model underpredicts the o/p ratio by up to an order of magnitude compared with a free-fall collapse model at gas densities n ≲ 106 cm−3 (the models are more consistent at higher densities). Consequently, it may not be unreasonable to expect variations in the o/p-H2D+ ratio across B2 as it evolves. Given this uncertainty in the o/p-H2D+ ratio, we limit discussion to N(ortho-H2D+), but note that the total N(H2D+) is likely greater than N(ortho-H2D+) by a factor of a few.

We show in Figure 9 the resulting N(ortho-H2D+) distribution across Oph B2. Including only pixels where S/N >5, we find the mean N(ortho-H2D+) = 1.4 × 1013 cm−2 with an rms variation of 0.7 × 1013 cm−2 over Oph B2. To this sensitivity limit, we find a minimum N(ortho-H2D+) = 4.1 × 1012 cm−2. The maximum N(ortho-H2D+) = 3.3 × 1013 cm−2 is found ∼30'' to the northwest of the 850 μm continuum peak. The largest ortho-H2D+ column densities are found in the western half of Oph B2, with moderate N(ortho-H2D+) values extending to the northeast.

Figure 9.

Figure 9. (a) Ortho-H2D+ column density in cm−2 (gray scale). N(ortho-H2D+) values have been divided by 1012. Gray contours show 850 μm continuum emission as in Figure 5. (b) Fractional ortho-H2D+ abundance X(ortho-H2D+) (gray scale). Values have been divided by 10−10.

Standard image High-resolution image

We next calculate X(ortho-H2D+) as described for N2D+ above and show in Figure 9(b) the X(ortho-H2D+) distribution across B2. Similar to the column density distribution, the largest H2D+ abundances, X(ortho-H2D+) ∼ (4–5) × 10−10, are found in the west. A local X(ortho-H2D+) minimum of ∼2 × 10−10 is found toward the central continuum emission peak, with similar abundances toward the northeast.

4.5. Using H2D+, N2H+ , and N2D+ to Determine TK

The molecular weight of H2D+ (mmol = 4.01 mH) is much less than that of N2H+ or N2D+ (29.01 mH and 30.02 mH, respectively). If we assume that the H2D+ 111–110, N2H+ 4–3, or N2D+ 3–2 transitions are excited in the same environment within Oph B2, we would expect each to trace the same core motions and be described by the same kinetic temperature TK. This co-location would imply that the non-thermal line widths for the two species should be equal, where σ2NT = σ2obs − σ2th. Given equal TK, the H2D+ thermal line width, σth = (kTk/mmol)1/2, will be measurably larger than that of N2H+ due to its smaller molecular weight. At TK = 12 K, for example, σth = 0.06 km s−1 for N2D+ and 0.16 km s−1 for H2D+. We can use this difference in thermal line widths to determine the kinetic gas temperature by equating the non-thermal line widths and solving for TK (e.g., in the case of N2H+):

Equation (7)

Mean temperatures of 9.2 ± 0.2 K, in good agreement with temperatures derived from NH3 alone, were derived in this manner toward a number of dense cores by Fuller & Myers (1993) using the molecular species HC3N and NH3. The largest source of error in the derived TK likely arises from the assumptions listed above rather than from the propagation of the small uncertainties on the observed line widths.

Overall, we expect the comparison of H2D+ and N2D+ to be most accurate, given the similar critical densities of the transitions, the fact that N2D+ must form from H2D+, and the similar distribution of integrated intensity (in contrast with the poor correlation between H2D+ and N2H+ 4–3 emission; see Figures 1(a) and 2). The different distributions of H2D+ and N2H+ 4–3 in Oph B2 suggest that these two species are not excited in the same environment in the core. Additionally, since our estimates of the N2H+ line widths are not accurate (given the unresolved hyperfine structure), we will focus only on N2D+ for this analysis.

We show in Figure 10 the temperatures calculated based on Equation (7) for H2D+ and N2D+ 3–2. Only pixels where S/N of both lines was >5 are shown. The derived TK values span a large range, from unphysical negative values (∼−11 K) to unrealistically large values (∼75 K). Using the mean Δv values of H2D+ and N2D+ from Table 2, we calculate a mean TK = 16 K, similar to that determined through NH3 analysis. Multiple pixels with TK ∼ 12 K are found surrounding the central continuum emission peak. Negative temperatures, however, are found associated with the continuum peak and associated clump B2-MM8.

Figure 10.

Figure 10. (a) TK determined using Equation (7) for H2D+ and N2D+ emission. Note that a negative TK was found for the blank pixel toward the submillimeter continuum emission peak because the observed H2D+ Δv > N2D+ Δv. Gray contours show 850 μm continuum emission as in Figure 5. (b) TK determined using Equation (7) for H2D+ and N2H+ 4–3 emission.

Standard image High-resolution image

From Equation (7), we calculate negative TK values when the observed H2D+ line widths are less than those of N2D+ 3–2. We show in Figure 11 (bottom) Δv derived for H2D+ and N2D+ emission (data values in 15'' pixels plotted), along with the spectra and relative fits toward B2-MM8 (top). Over most of B2, N2D+ line widths are narrower than H2D+ and thus TK > 0 K. The negative temperatures derive from two adjacent pixels toward B2-MM8 where the N2D+ Δv > Δv of H2D+. Figure 11 shows non-Gaussianity in the spectra of both species toward B2-MM8, which complicates the calculation of a single line width to describe the emission.

Figure 11.

Figure 11. Top left: H2D+ 111–110 spectrum toward B2-MM8 overlaid with a Gaussian fit. Top right: N2D+ 3–2 spectrum toward B2-MM8 overlaid with an HFS fit (note that no individual hyperfine components are visible due to their small separation in velocity and the relatively large line width). Bottom: returned H2D+ line widths Δv vs. N2D+ Δv. Each point represents a 15'' pixel. An equality relation is shown by the dotted line. Despite appearing wider, in most locations the N2D+ Δv is narrower than that of H2D+ due to the presence of overlapping hyperfine lines. The two points where N2D+ Δv > H2D+ Δv are located at and immediately adjacent to the 850 μm continuum peak and B2-MM8. At these locations, temperatures calculated via Equation (7) become negative and unphysical. Uncertainties shown are returned from the line fitting routines. Uncertainties in the line widths are shown.

Standard image High-resolution image

Why would we find greater N2D+ Δv than H2D+ Δv only toward the central continuum peak? One possibility is suggested by the spectra in Figure 11, where additional emission in the H2D+ spectrum is clearly seen blueshifted relative to the main Gaussian peak. For the H2D+ line, the best-fit Gaussian does not include the blue line shoulder. Slight blue asymmetry is also visible in the N2D+ line, but the distinction between the main component and the shoulder is less obvious. It is possible that the shoulder emission of N2D+ 3–2 is spread over a larger range in vLSR due to the contribution of multiple hyperfine components to the observed emission (each of which would contain emission at the central line velocity and at the shoulder velocity), broadening the line overall. The best fit to the N2D+ line thus cannot distinguish well between the main and shoulder emission, resulting in an artificially broad Δv.

5. DISCUSSION

5.1. Line Widths and Density

The mean line widths listed in Table 2 show that on average the observed Δv for a given molecular line decreases with increasing critical line density, suggesting that the gas in B2 becomes more quiescent at higher densities (omitting N2H+ 4–3, where Δv is an overestimate). For an average gas temperature TK = 15 K, the resulting non-thermal line contribution decreases from σNT = 0.36 km s−1 for NH3 (1,1) emission to σNT = 0.22 km s−1 for N2D+ 3–2 emission (from ncr ∼ 103–104 cm−3 to ncr ∼ 8 × 105 cm−3, respectively). Comparing with the thermal sound speed at 15 K, where $c_s = \sqrt{kT_K/(\mu m_{{\rm H}})}$, we find an average σNT / cs = 1.6 for NH3 and σNT / cs = 1.1 for N2D+ 3–2 emission. This comparison shows that despite the decrease in non-thermal motions, the non-thermal motions are equal in magnitude with the sound speed, even at high density. A lower gas temperature would result in even larger magnitudes of σNT, while significantly higher temperatures are unlikely.

This decrease in non-thermal motions at higher densities has been observed in starless cores (e.g., Pon et al. 2009; Tafalla et al. 2004; Lada et al. 2003). Frequently in starless cores, however, the narrowest observed line widths are consistent with being due to nearly pure thermal motions, in contrast to the results found here. As σNT values decrease, the ability for turbulent pressure to support the core against collapse is reduced, which impacts the stability of the core against its self-gravity. The narrow σNT widths found in starless cores are often explained as the result of the dissipation of turbulent motions in the core center (Goodman et al. 1998). If the non-thermal line widths in B2 are due to turbulent motions, the corresponding turbulent pressure (PNT = mnσ2NT) is equal to or greater than the thermal pressure (PT = nkT).

In Paper I, we discussed the source of wide NH3 line widths in Oph B and concluded that if the relatively large σNT values are caused by turbulent motions, they could not be primordial (i.e., inherited from the parent cloud). This is because the damping timescale of such motions (with no external driving force) is approximately the dynamical time (Mac Low & Klessen 2004), and the existence of protostars associated with B2 implies that B2 has likely existed, with high density gas, for at least a dynamical time.

The wide lines in Oph B2 may, however, be due to more organized motions such as infall or outflow. None of the observed molecular lines are optically thick, and we therefore do not see the self-absorbed, asymmetric line profiles frequently used to infer the infall of gas within a core. Infall signatures were observed in CS, H2CO, or HCO+ emission toward several locations in eastern B2 by André et al. (2007), but not in central or western B2. Gurney et al. (2008) observed complicated CO (and CO isotopologues) line structures toward the B2 continuum peak but were unable to distinguish infall conclusively. The analysis of line profiles is complicated by a substantial outflow discovered in CO emission and emanating from one of the two embedded YSOs in B2 (Kamazaki et al. 2003) which extends over 10' (∼0.4 pc) in size.9 Upcoming results of CO, 13CO, and C18O 3–2 observations in Ophiuchus (G. White et al. 2010, in preparation) from the JCMT Gould Belt Legacy Survey (GBLS; Ward-Thompson et al. 2007) will be better able to study the infall and outflow motions in Oph B. In conjunction with these data, the process of how the outflow motions affect the dense gas may be studied. There is clear evidence, however, for protostellar influence on the N2D+ line widths shown in Figure 5(b), as discussed in Section 4.2. Additionally, the blue shoulder emission seen in N2H+ and N2D+ east of E32 and E33 coincides with the eastern CO lobe of the outflow and is further evidence for the significant impact of the outflow on the dense gas.

We also noted in Section 4.2 that there is an offset in vLSR of ∼0.2 km s−1 between N2H+ 4–3 and H2D+ emission toward B2-MM8 and the nearby embedded protostar Elias 32, with the N2H+ emission redshifted relative to the H2D+ emission. Over the rest of Oph B2, the correlation between the vLSR of the two lines is good, suggesting that the offset is real and localized. Comparison with N2D+ emission at the same location shows that the vLSR of the H2D+ and N2D+ agree well, and it is consequently the N2H+ emission which is offset in velocity from the rest of the core gas. N2H+ 4–3 has a critical density approximately an order of magnitude greater than the H2D+ and N2D+ lines discussed here. The offset in vLSR may thus be probing a shift in the gas motions at extremely high densities in the core interior.

5.2. Trends in the Deuterium Fractionation

5.2.1. The N2D+/N2H+ Ratio

The deuterium fractionation, RD, can be defined as the ratio between the column densities of a deuterated molecule and its hydrogen-bearing counterpart. Here, we define RD = N(N2D+)/N(N2H+) and calculate RD in Oph B2 from N(N2D+) calculated in Section 4.4.1 and N(N2H+) calculated in Paper II. Figure 7(b) shows RD across B2. The distribution is similar to the N2D+ column density in Figure 7(a). The largest values of RD, to a maximum RD = 0.16, are found toward northeast B2, while toward the smaller scale maxima, we find RD ∼ 0.1 or less. We estimate that we are sensitive to RD ∼ 0.01–0.015 based on the rms noise in the N2D+ integrated intensity map (3σ ∼ 0.2 K km s−1 in the off-B2 pixels), and the mean N2H+ column density over Oph B2, 〈N(N2H+)〉 = 5.9 × 1012 cm−2, calculated in Paper II. We find a moderate mean RD = 0.03 with an rms variation of 0.01 toward B2, comparable to results in Ori B9 toward two N2H+ emission peaks and a protostellar source (RD = 0.03–0.04, Miettinen et al. 2009). While Pagani et al. (2009a, 2009b) find substantially greater central RD values through modeling of the starless core L183 (RD ∼ 0.7 at the core center) which decrease to ∼0.06 at ∼5000 AU, it is not clear how to compare their model of RD as a function of core radius to our observed column density ratio. Fontani et al. (2006) found an average RD ∼ 0.01 toward the high-mass star-forming region IRAS 05345+3157, which at high resolution resolved into two N2D+ condensations each with RD = 0.11 (Fontani et al. 2008). In a study of starless cores, Crapsi et al. (2005) find a range in RD between a lower limit of 0.02 and a maximum of 0.44 (which was found toward the Oph D core). Crapsi et al. also note that RD values >0.1 are generally only found for cores with N(N2H+)>1013 cm−2, which are approximately the maximum N(N2H+) in Oph B2. A similar range in RD was found toward protostellar, Class 0 sources (Emprechtinger et al. 2009). On average, then, our results agree with previous results. Given our map of Oph B2, however, we can also probe the variation of RD across the core.

In Figure 12, we show RD as a function of increasing N(H2) (Figure 12(a)), and also as a function of increasing distance (in projection) from the nearest embedded protostar (Figure 12(b)). Each data point represents an 18'' pixel (to match the beam FWHM), and pixels with S/N <5 in either N2H+ or N2D+ emission are omitted. No clear trend is present in the deuterium fractionation with N(H2). We find, however, that RD increases at greater projected distances from the embedded sources, as might be expected from the N2D+ integrated intensity distribution. We fit a linear trend to the data, where RD = (−0.007 ± 0.002) + (0.37 ± 0.06)d [pc], but note that there is large scatter in the relation. An analysis of the individual N2H+ and N2D+ abundances shows that the increasing RD trend with projected protostellar distance is dominated by a similar increase in X(N2D+) with projected protostellar distance, while X(N2H+) is approximately constant over the range plotted, shown in Figure 12(d).

Figure 12.

Figure 12. Variation in RD, X(N2H+), X(ortho-H2D+), and TK with H2 column density and the projected distance to the nearest embedded protostar in Oph B2. In each plot, points shown represent pixels matched to the FWHM beam of the respective data sets (15'' for TK and H2D+, and 18'' for N2H+ and N2D+). Typical uncertainties are shown for a single pixel in each plot. An uncertainty of ∼20%–30% in the X(ortho-H2D+) values is not shown. Only pixels with S/N >5 are plotted. In (a), no trend is found in RD vs. log N(H2). The RD data are, however, consistent with an increasing deuterium fractionation with distance from an embedded source (b), albeit with significant scatter. As discussed in Paper II, a strong correlation is seen in X(N2H+) vs. log N(H2) (c), but no trend is seen with respect to projected protostellar distance (d). X(ortho-H2D+) shows no trend with log N(H2) (e). With respect to projected protostellar distance, the X(ortho-H2D+) plot shows that the data do not scatter about a central point (f), but instead show a trend of increasing X(H2D+) to d ∼ 0.04 pc, and a smaller, constant H2D+ abundance at further distances. In both (e) and (f), the points identified with squares represent pixels in the northeast portion of the B core. A linear fit to the data, omitting the highlighted pixels, is shown by the dotted line. A linear fit to the data is shown by the dashed line. No clear correlation is found in TK vs. log N(H2) (g) or projected protostellar distance (h).

Standard image High-resolution image

5.2.2. H2D+

We next look at the distribution of the ortho-H2D+ abundance over Oph B2. In Figure 12, we show X(ortho-H2D+) as a function of N(H2) (Figure 12(e)) and also as a function of increasing projected protostellar distance (Figure 12(f)). We show data in 15'' pixels and omit pixels with S/N <5. We find no trend in the fractional ortho-H2D+ abundance with H2 column density (an increasing trend is found, but not shown, in the ortho-H2D+ column density, N(ortho-H2D+), with N(H2)). The X(ortho-H2D+) versus N(H2) relationship is not entirely a scatter plot, as the ortho-H2D+ abundance tends to increase with projected protostellar distance to a maximum X(ortho-H2D+) = 4.9 × 10−10 at d = 0.035 pc, with X(ortho-H2D+) ∼ 1.8 × 10−10 at larger distances. All data points at d > 0.035 pc correspond to the emission found in northeast B2, while the rest show emission closer to the central continuum emission peak. If we fit a linear relationship to the data, omitting the northeast B2 points, we find X(ortho-H2D+) × 10−10 = (1.2 ± 0.5) + (79 ± 23)d [pc].

There is also no clear trend in N(ortho-H2D+) with RD where both H2D+ and N2D+ are strongly detected across Oph B2, although both species are found with relatively large column densities north and west of the continuum peak. Given the small-scale structure found in the N(N2D+) distribution in B2, the spatial resolution and sampling of our H2D+ observations may not be high enough to discern a correlation. We note, however, that Caselli et al. (2008) also found no clear correlation between RD and N(ortho-H2D+) in their sample of starless and protostellar cores (with the caveat that some RD values were derived using column density ratios of deuterated NH3 and deuterated H2CO with their undeuterated counterparts rather than N(N2D+)/N(N2H+)). In particular, high values of RD and relatively low N(ortho-H2D+) were found toward two objects in Ophiuchus (Oph D and 16293E).

5.3. What is Affecting the Deuterium Fractionation in Oph B2?

The largest deuterium fractionation values in Oph B2 are not found toward the continuum emission peak but instead avoid the continuum peak and nearby protostars. The results of the previous section indicate that the distance to the nearest protostar is the dominant parameter impacting the deuterium fractionation in Oph B2. We next discuss several mechanisms by which the embedded sources are most likely to impact the deuterium fractionation distribution in B2, including temperature variation in the gas, liberation of CO from dust grains, the impact of protostellar outflows, and possible increase in the ionization fraction due to X-rays from the protostars.

5.3.1. Temperature and CO Depletion

Temperature can have a substantial impact on the deuterium fractionation in a dense core. As stated in Section 4.4.1, we expect the gas and dust to be well coupled at the densities of Oph B2 such that Td = Tgas but specify in the following which of the gas or dust temperature is important in the following mechanisms. First, at Tgas ≳ 20 K, the H2D+ reaction in Equation (1) can proceed both forward and backward, resulting in no net increase in H2D+, and consequently other deuterated species formed via reactions with H2D+. Second, the CO which was deposited (adsorbed) as an icy mantle onto dust grains while at low temperatures will evaporate (desorb) back into the gas phase if the dust grains are heated, providing a destruction mechanism for H2D+ and again interrupting the deuteration chain.

The CO depletion factor, fD(CO), is defined as the canonical CO abundance, Xcan(CO), divided by the CO abundance, X(CO) = N(CO)/N(H2), in the observed region:

Equation (8)

such that fD values are low where observed CO abundances are high. Low CO depletion factors can dramatically lower the H2D+ abundance at gas temperatures T < 15 K (Caselli et al. 2008). Currently, no estimates of fD exist in B2 due to a lack of data at spatial resolutions matching the available continuum data. The CO observations of the JCMT GBLS should enable the first calculations of fD at complementary spatial resolution and allow further analysis of the relationship between CO and the deuterium distribution in B2. The extended distribution of H2D+ and N2D+ across Oph B2 implies, however, that extensive depletion of CO has occurred. At densities of n = 106 cm−3, CO adsorption and desorption occur equally frequently when Td = 18 K (Visser et al. 2009). At Td > 18 K, CO will thus again become significantly more abundant in the gas phase, while at Tgas > 20 K, enhanced deuterium fractionation via Equation (1) will cease to operate.

An obvious potential source of heating is the presence of embedded protostars within Oph B2. In a one-dimensional dynamic/chemical model of an isolated dense core which collapses to form a protostar, Aikawa et al. (2008, see their Figure 2) find that temperatures greater than 18 K can be found at radii ∼5000 AU (∼0.02 pc) from the protostar at ∼105 yr after its formation. As a function of total luminosity of the core, Lcore, however, their model predicts a CO sublimation radius of ≲1000 AU (0.005 pc) for Lcore ∼ 0.5 − 1 L. In a more specific example, Stamatellos et al. (2007) model the heating effect of an embedded Class I protostar (Lbol = 10 L) on the surrounding gas in the Oph A core and find only a small temperature increase (≲5 K), limited to the immediate vicinity of the protostar, in agreement with Aikawa et al. (2008).

Both YSOs in central B2 have been classified as embedded (Jørgensen et al. 2008), Class I sources (Enoch et al. 2007). The estimated bolometric luminosities, Lbol ∼ 0.5 L and ∼1 L for E32 and E33, respectively (Enoch et al. 2007; van Kempen et al. 2009), are much smaller than used in the heating model by Stamatellos et al. (2007). The bolometric temperatures, Tbol ∼ 300 K and ∼500 K, suggest that both are in the later stages of the Class I phase, as do their negative infrared spectral indices (αIR = −0.03 and −0.12). Sources with −0.3 < αIR < 0 are sometimes classified as 'flat-spectrum' sources and are thought to be an intermediate stage between Class I and Class II. Crapsi et al. (2008) suggest that most or all flat-spectrum sources could be Class II objects seen edge-on. Based on these results, E32 and E33 are not likely to be deeply embedded within Oph B and have ages ∼5 × 105 yr or greater based on lifetime estimates of the different protostellar stages (Evans et al. 2009).

The low luminosities of E32 and E33 suggest that they should not have a large impact on the temperature of B2, in agreement with our temperature results in Paper I which showed very little variation in the gas temperature over Oph B. Figure 12 shows that we find no trend in the kinetic temperature TK, determined using NH3 inversion transitions in Paper I, with protostellar distance above the ∼1–2 K uncertainties in the data. The derived temperatures are near the required values, however, and the NH3 temperatures may not be indicative of the conditions of the denser gas where we expect most of the deuterated species to exist. Depletion of H2D+ and N2D+ via CO sublimation would thus require an increasing temperature gradient toward higher densities near the protostars, which is not unreasonable.

Since CO also destroys N2H+, significant evaporation of CO from dust grains should also be observable in the N2H+ distribution. Accordingly, in Figure 1(b) we see a decrease in the integrated N2H+ 1–0 intensity toward E33. No good fit to the N2H+ spectra was found at this location, however, due to complicated line structures, such that we are not able to determine the N2H+ column density or abundance. Figure 12(d) shows no trend on larger scales of X(N2H+) with protostellar proximity. The N2H+ 4–3 distribution shows that N2H+ remains in the gas toward E33, but an analysis of the excitation of the line would be needed to determine the N2H+ column density from these data.

In summary, heating by an embedded protostar can produce slightly greater gas temperatures and evaporate CO from the surface of dust grains, which would decrease the deuterium fractionation in the gas. Since we do not see evidence of a decrease in X(N2H+) toward the protostars, the evaporation of CO from dust grains is likely not the major mechanism for the decrease in abundance of deuterated species. While we also do not see evidence of increased gas temperatures near the protostars in the NH3 data, these temperatures reflect average values along the line of sight, and do not rule out increased temperatures at higher densities near the protostars. Given the strong dependence of increased deuterium fractionation on temperature, we expect this mechanism is likely important in B2.

5.3.2. The Oph B2 Outflow

The deuterium fractionation may be impacted by the known, extensive CO outflow in B2 described in Section 5.1. It appears that the dense gas in eastern Oph B2 is affected by the outflow, given the larger line widths seen toward the protostars in high density tracers, changes in vLSR, and the blue shoulder emission which is coincident with the eastern lobe of the outflow (Kamazaki et al. 2003; although large line widths could also indicate the presence of infall motions). We do not find evidence of outflow interaction with the dense gas in northwest Oph B2, suggesting that the outflow may be in front of or behind the bulk of the core gas. Very localized heating of the surrounding gas by the protostar could evaporate CO from the dust grains. The outflow, then containing entrained CO, could mix the CO with the dense core gas, destroying the deuterated species without an increase in the gas temperature on larger scales. Alternatively, shocks associated with the outflow could raise gas temperatures and reintroduce enough CO to the gas phase to affect the deuterium fractionation. Further analysis of the outflow, including observations of shock tracers, is needed to determine whether this scenario is reasonable and can examine the impact of low-mass protostellar outflows on dense gas chemistry. As discussed above, however, we do not find evidence for variation in the N2H+ abundance as would be expected if destruction due to gas-phase CO is the dominant mechanism for decreasing the abundance of deuterated species relative to their non-deuterated counterparts.

5.3.3. Ionization Fraction and Electron Abundance xe

Variations in the ionization fraction in the dense gas can also affect the deuterium fractionation, in particular in regions of high CO depletion. The electron abundance, $x_e = n_e/n_{\rm H_2 }$, is the ratio of the electron number density (ne) with the number density of molecular hydrogen, $n_{\rm H_2}$, and is equivalent to the ionization fraction. The deuterium fraction can be reduced by an increase in xe, which in turn increases the rate of dissociative recombination (Caselli et al. 2008). The abundances of deuterated molecules can thus be used as a probe of xe in dense cores. At high extinction toward starless cores, the creation of ions is expected to be dominated by the cosmic ray flux, with $x_e \propto n_{\rm H_2}^{-1/2}$ (McKee 1989).

Studies investigating xe in dense cores using the [DCO+]/[HCO+] ratio have found average line-of-sight xe ∼ 10−7 (Bergin & Tafalla 2007, and references therein). In regions of high depletion, the column density ratio of N2D+ to N2H+ is expected to be a better probe of xe than [DCO+]/[HCO+]. Since H+3 is expected to be the major molecular ion in cores strongly depleted in heavy elements, we can use observations of H2D+, in conjunction with RD, to set limits on xe assuming that the numbers of positive and negative charges are approximately equal (i.e., the core gas is neutral). Given the abundances of N2H+, N2D+, and H2D+ in B2, we can thus derive a lower limit on xe following Miettinen et al. (2009):

Equation (9)

where we neglect multiply deuterated forms of H+3. As discussed in Section 4.4.2, predictions of the o/p-H2D+ ratio vary from ∼0.1 to ≳1, and we use a moderate o/p-H2D+ = 0.5 to estimate X(H2D+) from X(ortho-H2D+). In a simple, steady state analytical model, Crapsi et al. (2004) relate RD (where RD = N(N2D+) / N(N2H+)) to the relative H2D+–H+3 abundance, r = [H2D+]/[H+3], and also neglect multiply deuterated forms of H+3 such that RDr/(3 + 2r). In an updated result, Miettinen et al. (2009) find RD ≈ (r + 2r2)/(3 + 2r + r2). Using the Miettinen et al. relation, we then find a mean r = 0.07 across Oph B2 and a lower limit xe > 1.5 × 10−8 through Equation (9). We find small variations in the lower limit across B2, with a larger limit xe > 2.7 × 10−8 toward Elias 32 and higher values in general near the 850 μm continuum peak. Detailed modeling, in conjunction with a measurement of the CO depletion factor in B2, is needed to improve this analysis of xe beyond a lower limit only.

Low mass protostars emit X-rays, which could potentially increase the local ionization fraction and consequently decrease locally the deuterium fractionation of gas in Oph B2. In their study of the X-ray emission from young protostars in the central Ophiuchus region with ROSAT, Casanova et al. (1995) detected both Oph B2 protostars in the 1–2.4 keV range, finding luminosities (adjusted for extinction) Lx = 30.7 erg s−1 and Lx = 29.2 erg s−1 for E32 and E33, respectively. We can then estimate the X-ray ionization rate due to the protostars, ζx, at a distance r following Silk & Norman (1983, Equation (2)). From Figure 12, we find trends in RD and X(ortho-H2D+) to projected protostellar distances of order of a few ×0.01 pc. Assuming a constant, uniform density n = 105 cm−3, at r = 0.02 pc we find ζx ∼ 9 × 10−17 s−1, and at r = 0.04 pc we find ζx ∼ 8 × 10−18 s−1. The fractional ionization based on these rates, again following Silk & Norman, is xe ∼ 9 × 10−8 and xe ∼ 3 × 10−8 at r = 0.02 pc and r = 0.04 pc, consistent with our lower limits calculated above. Clumpy, rather than uniform gas could allow X-rays to penetrate further into the core, potentially increasing the fractional ionization to greater distances. The standard ionization rate due to cosmic rays alone ζCR = 1.3 × 10−17 s−1 (Herbst & Klemperer 1973), leading to a fractional ionization xe ∼ 4 × 10−8 for n = 105 cm−3 (McKee 1989). Our simple calculation shows that at distances comparable to where we find variation in the deuterium fractionation in Oph B2, the protostellar X-ray ionization rate and resulting fractional ionization may be comparable to or greater than expected due to cosmic rays alone. This result suggests that X-ray ionization is a viable mechanism to depress the deuterium fractionation close to embedded protostars. Further work is needed to directly measure the fractional ionization of dense gas near protostars and to quantify the effect of a greater ionization rate on the chemical network of deuterated species in high density gas.

5.3.4. Other Possibilities

While Figure 12 suggests that the distribution of enhanced deuterium fractionation of H+3 and N2H+ in Oph B2 is tied to the presence of the Class I protostars associated with the core, other factors may influence the deuteration chemistry in this complex environment. For completeness, we briefly mention two other possibilities. First, line-of-sight inhomogeneities in Oph B2 would impact our estimate of the deuterium fractionation, as our analysis assumes constant excitation and abundance conditions along the line of sight. Detailed physical and chemical modeling may be useful in probing the effects of possible inhomogeneities but is beyond the scope of this paper due to the complexity of Oph B2. Second, differences in the evolutionary timescale may be important in Oph B2 and in other multiple-star-forming objects. While several protostars have already formed out of the core gas (particularly in the south), there remain multiple dense, starless clumps within B2 which may yet form stars. The timescale on which different parts of the core condense and collapse may impact the dense gas chemistry and produce variations in the deuterium fractionation across Oph B2.

In summary, there a number of different mechanisms by which the deuterium fractionation of N2H+ and H+3 may be decreased near young protostars. We feel that a local increase in the gas temperature is the most likely explanation of this trend, but our simple calculation of the expected increase in the ionization fraction above the standard cosmic ray value is intriguing and worth further study.

6. SUMMARY

In this paper, we have presented the results of N2D+ 3–2, N2H+ 4–3, and ortho-H2D+ 111–110 mapping of the cluster-forming Ophiuchus B core. In general, our results show that care must be taken while using deuterated species as a probe of the physical conditions of dense gas conditions in star-forming regions. We summarize in detail our results below.

  • 1.  
    Significant emission from the deuterated species N2D+ and H2D+ extends over several square arcminutes in B2. In particular, the extent of the H2D+ emission is the largest yet mapped. Both species' integrated intensity distributions are offset from the continuum emission and avoid entirely parts of B2 near embedded protostars and the peak of continuum emission from cold dust.
  • 2.  
    Through fitting of the spectral lines, we find greater N2D+ line widths near protostars embedded in B2, suggestive of protostellar influence on the dense gas. Complicated line structures are found, including blue line shoulders, in all three tracers. We estimate the non-thermal line widths for the species observed here and compare with NH3 and N2H+ results from Paper I and Paper II. In B2, non-thermal line widths decrease with increasing critical density of the molecular tracer but remain transonic or only slightly subsonic even at high densities. This result is in contrast with the nearly thermal line widths found in many isolated cores.
  • 3.  
    We test a method to estimate the gas temperature by equating the non-thermal line widths of N2D+ and H2D+ and solving for TK. We find both realistic (TK ∼ 11 K) and unrealistic (TK < 0 K and TK ≳ 25 K) values in B2. Complicated line structure in B2 likely affects this analysis, but we expect good results can be found in regions with simple velocity structure.
  • 4.  
    We find RD = N(N2D+)/N(N2H+) ∼ 0.03 on average where N2D+ is detected. Small-scale features of enhanced RD are seen in B2, but most of these do not correlate with previously identified continuum clumps or N2H+ clumps. This RD is consistent with previous results in protostellar cores and is also within the lower range of values found for starless cores.
  • 5.  
    The average ortho-H2D+ abundance, X(ortho-H2D+) ∼ 3 × 10−10, given an assumed excitation temperature of 12 K. An accurate measure of the ortho- to para-H2D+ ratio is needed to convert to total H2D+ abundance and has been found in previous studies to vary between 0.1 and >1. Future observations of para-H2D+ transitions, if possible, are needed to constrain this value in dense cores.
  • 6.  
    An anti-correlation is found with RD and X(ortho-H2D+) and projected distance to the nearest embedded protostar in B2. This relation indicates that the embedded protostars in B2 are affecting the internal chemistry. Possible mechanisms are through gas heating, evaporation of CO off dust grains, CO mixing caused by the B2 outflow, or by increasing the local ionization fraction. Gas temperatures determined through NH3 analysis in Paper I are <18–20 K required to evaporate CO and reduce deuterium fractionation, but NH3 may not be a good tracer of conditions in the high density gas in Oph B2. We calculate that the X-ray flux of the embedded protostars in B2 could increase the fractional ionization of the core gas beyond that expected solely due to cosmic ray ionization to distances ∼0.02–0.03 pc, similar to the distances over which we find variation in the deuterium fractionation. We suggest that a temperature increase is the dominant parameter in this trend, but further research on the effect of young stars and their outflows on both the local ionization fraction and the chemistry of deuterated species in protostellar cores is needed.
  • 7.  
    We use the observed RD and X(ortho-H2D+) to estimate a lower limit on the ionization fraction in B2 and find x(e)>1.5 × 10−8 given an assumed ortho- to para-H2D+ ratio of 0.5.

We thank the anonymous referee and E. Feigelson for comments, which improved the paper. We also thank H. Kirk for providing SCUBA maps of the observed regions and H. Weisemeyer for help with the IRAM observations. The National Radio Astronomy Observatory is a facility of the National Science Foundation operated under cooperative agreement by Associated Universities, Inc. 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. This research is supported in part by the National Science Foundation under grant number 0708158 (T.L.B.).

Facilities: JCMT - James Clerk Maxwell Telescope, IRAM:30 m - Institute de Radioastronomie Millimetrique 30 meter telescope

Footnotes

Please wait… references are loading.
10.1088/0004-637X/718/2/666