Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
See discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/51966952 A (sub)millimetre study of dense cores in Orion B9 Article in Astronomy and Astrophysics · December 2011 DOI: 10.1051/0004-6361/201117849 · Source: arXiv CITATIONS READS 18 41 4 authors, including: Jorma Harju University of Helsinki 86 PUBLICATIONS 780 CITATIONS SEE PROFILE All content following this page was uploaded by Jorma Harju on 08 December 2016. The user has requested enhancement of the downloaded file. Astronomy & Astrophysics manuscript no. miettinen˙et˙al˙2011b December 22, 2011 c ESO 2011 A (sub)millimetre study of dense cores in Orion B9⋆,⋆⋆ O. Miettinen1 , J. Harju2,1 , L. K. Haikala2,1 , and M. Juvela1 1 2 Department of Physics, P.O. Box 64, FI-00014 University of Helsinki, Finland e-mail: oskari.miettinen@helsinki.fi Finnish Centre for Astronomy with ESO (FINCA), University of Turku, Väisäläntie 20, FI-21500 Piikkiö, Finland arXiv:1112.5053v1 [astro-ph.GA] 21 Dec 2011 Received ; accepted ABSTRACT Context. Studies of dense molecular-cloud cores at (sub)millimetre wavelengths are needed to understand the early stages of star formation. Aims. We aim to further constrain the properties and evolutionary stages of dense cores in Orion B9. The prime objective of this study is to examine the dust emission of the cores near the peak of their spectral energy distributions, and to determine the degrees of CO depletion, deuterium fractionation, and ionisation. Methods. The central part of Orion B9 was mapped at 350 µm with APEX/SABOCA. A sample of nine cores in the region were observed in C17 O(2 − 1), H13 CO+ (4 − 3) (towards 3 sources), DCO+ (4 − 3), N2 H+ (3 − 2), and N2 D+ (3 − 2) with APEX/SHFI. These data are used in conjunction with our previous APEX/LABOCA 870-µm dust continuum data. Results. All the LABOCA cores in the region covered by our SABOCA map were detected at 350 µm. The strongest 350 µm emission is seen towards the Class 0 candidate SMM 3. Many of the LABOCA cores show evidence of substructure in the higher-resolution SABOCA image. In particular, we report on the discovery of multiple very low-mass condensations in the prestellar core SMM 6. Based on the 350-to-870 µm flux density ratios, we determine dust temperatures of T dust ≃ 7.9 − 10.8 K, and dust emissivity indices of β ∼ 0.5 − 1.8. The CO depletion factors are in the range fD ∼ 1.6 − 10.8. The degree of deuteration in N2 H+ is ≃ 0.04 − 0.99, where the highest value (seen towards the prestellar core SMM 1) is, to our knowledge, the most extreme level of N2 H+ deuteration reported so far. The level of HCO+ deuteration is about 1–2%. The fractional ionisation and cosmic-ray ionisation rate of H2 could be determined only towards two sources with the lower limits of ∼ 2 − 6 × 10−8 and ∼ 2.6 × 10−17 − 4.8 × 10−16 s−1 , respectively. We also detected D2 CO towards two sources. Conclusions. The detected protostellar cores are classified as Class 0 objects, in agreement with our previous SED results. The detection of subcondensations within SMM 6 shows that core fragmentation can already take place during the prestellar phase. The origin of this substructure is likely caused by thermal Jeans fragmentation of the elongated parent core. Varying levels of fD and deuteration among the cores suggest that they are evolving chemically at different rates. A low fD value and the presence of gas-phase D2 CO in SMM 1 suggest that the core chemistry is affected by the nearby outflow. The very high N2 H+ deuteration in SMM 1 is likely to be remnant of the earlier CO-depleted phase. Key words. Astrochemistry - Stars: formation - ISM: abundances - ISM: clouds - ISM: molecules - Radio lines: ISM - Submillimeter: ISM 1. Introduction New stars in our Galaxy form predominantly in the so-called giant molecular clouds (GMCs). The nearest GMC to the Sun is the Orion molecular cloud complex1. This cloud complex contains a number of star-forming regions, one example being the Orion B9 region in the central part of Orion B molecular cloud. Orion B9 represents a very early stage of star formation which is manifest in the fact that most “dense cores” in the cloud are either prestellar or contain Class 0 protostars. These objects are Send offprint requests to: O. Miettinen ⋆ This publication is based on data acquired with the Atacama Pathfinder EXperiment (APEX) under programmes 079.F-9313A, 084.F-9304A, and 084.F-9312A. APEX is a collaboration between the Max-Planck-Institut für Radioastronomie, the European Southern Observatory, and the Onsala Space Observatory. ⋆⋆ Our SABOCA map shown in Fig. 2 is available electronically in FITS format at the CDS. 1 In this paper we adopt a distance of 450 pc to the Orion GMC (Genzel & Stutzki 1989). The actual distance may be somewhat smaller as, for example, Menten et al. (2007) determined the distance to the Orion Nebula to be 414 ± 7 pc. cold, and their study requires observations at (sub)millimetre wavelengths. Using the LABOCA bolometer array on APEX (Atacama Pathfinder EXperiment), we mapped that region at 870 µm (Miettinen et al. 2009; hereafter Paper I). The dust continuum mapping resulted in the discovery of twelve dense cores of which four were found to be associated with IRAS point sources, and eight of them appeared previously unknown sources. Of the newly discovered cores, two (we called SMM 3 and 4) were found to be associated with Spitzer 24- and 70-µm sources, and were tentatively classified as Class-0 protostellar candidates. Six cores showed no signs of embedded infrared emission, and were thus classified as starless. Figure 1 shows the 870 µm emission as contours overlaid on the Spitzer/MIPS 24-µm image of the central part of Orion B92 . In Paper I, we also de2 Comparison of our LABOCA map to the new SABOCA map together with Spitzer images showed that the LABOCA observations were not well-pointed. The SABOCA peaks appeared to systematically lie southeast from the LABOCA peak positions. The difference in angular resolution between SABOCA and LABOCA data did not appear to be the cause of the offset in peak positions; by smoothing 1 Miettinen et al.: Dense cores in Orion B9 rived the degree of deuteration, or the N(N2 D+ )/N(N2 H+ ) column density ratio, towards selected positions and found the values in the range 0.03–0.04, comparable to those seen in other low-mass star-forming regions (e.g., Crapsi et al. 2005; Emprechtinger et al. 2009; Friesen et al. 2010b). Taking advantage of the H2 D+ data from Harju et al. (2006), the ionisation degree and the cosmic-ray ionisation rate of H2 towards the same target positions were estimated to be x(e) ∼ 10−7 and ζH2 ∼ 1 − 2 × 10−16 s−1 , respectively. The physical properties of dense cores in Orion B9 were studied further by Miettinen et al. (2010; hereafter Paper II). Using the observations of the NH3 (1, 1) and (2, 2) inversion transitions performed with the Effelsberg 100-m telescope, we determined the gas kinetic temperature, kinematical properties, and dynamical state of the cores. The gas kinetic temperature of the cores were found to be T kin ∼ 9.4 − 13.9 K, and the internal non-thermal motions in the cores appeared to be subsonic, or at most transonic. The virial-parameter analysis showed that the starless cores in the region are likely to be gravitationally bound, and thus prestellar objects. Interestingly, some of the cores were found to have significantly (by ∼ 5 − 7.5 km s−1 ) lower radial velocity than the systemic velocity of the region (vLSR ∼ 9 km s−1 ). This suggests that they belong to the “low-velocity part” of Orion B which is likely to originate from the feedback from the massive stars of the nearby Ori OB 1b association. In this paper, we present the results of our new APEX observations of the Orion B9 region. These include the dust continuum mapping with the SABOCA (Submillimetre APEX Bolometer Camera) bolometer array at 350 µm, and observations of the molecular-line transitions C17 O(2 − 1), H13 CO+ (4 − 3), DCO+ (4 − 3), N2 H+ (3 − 2), and N2 D+ (3 − 2). The SABOCA 350-µm observations allow us to probe the peak of the core’s spectral energy distributions (SEDs). We also aim to study the degrees of CO depletion and deuterium fractionation, and the fractional ionisation in the cores. These are useful parameters to further constrain the physical and chemical properties of the sources, and their evolutionary stages. This paper is structured as follows. The observations and data-reduction procedures are described in Sect. 2. The direct observational results are presented in Sect. 3. In Sect. 4, we describe the analysis and present the results of the physical and chemical properties of the cores. Discussion of our results is presented in Sect. 5, and in Sect. 6, we summarise the main conclusions of this study. 2. Observations and data reduction 2.1. Submillimetre dust continuum observations The central 18.′ 0 × 14.′ 5 part of Orion B9 was mapped with SABOCA (Siringo et al. 2010) on the APEX 12-m telescope (Güsten et al. 2006) at Llano de Chajnantor (Chilean Andes). SABOCA is a 37-channel on-sky bolometer array operating at 350 µm, with a nominal resolution of ∼ 7.′′ 5 (HPBW). The effective field of view of the array is 1.′ 5. The SABOCA passband the SABOCA map to correspond the resolution of our LABOCA data, and regridding the maps onto the same grid, the offsets still remained. We used the brightest SABOCA 350-µm source in the region, SMM3, as a reference to adjust the pointing of our LABOCA map. The pointing of the LABOCA map was shifted by (∆α, ∆δ) = (+1.′′ 72, −9.′′ 97). Unfortunately, the target positions of our molecular-line observations, which were chosen to be the LABOCA peak positions from our previous map, are now slightly offset from the dust maxima (but still within the beam size for most line observations). 2 has an equivalent width of about 120 GHz centred on an effective frequency of 852 GHz. The observations were carried out on 5 October and 10 November 2009. The atmospheric zenith opacity at 350 µm was measured using the sky-dip method, and was found to be in the range τz350 µm = 0.8 − 1.1. The amount of precipitable water vapour (PWV) was in the range 0.4–0.6 mm. The telescope pointing and focus checks were made at regular intervals using the planets Mars, Jupiter, and Uranus, and several secondary calibrators (e.g., CRL618, N2071IR, and CW Leo). The absolute calibration uncertainty for SABOCA is 25–30%. Mapping was performed using a ’fast-scanning’ method without chopping the secondary mirror (Reichertz et al. 2001). The mosaic was constructed by combining 147 individual fast-scanning (typically 1.′ 2 s−1 ) maps. The total integration time spent on the area was 9.7 hr. Data reduction was done with the CRUSH-2 (Comprehensive Reduction Utility for SHARC-2) (version 2.03-2) software package (Kovács 2008), which includes calibration data covering the period of our SABOCA observations. We used the pipeline iterations with specifying the ’deep’ option, which is appropriate for point-like sources. To improve the image appearance and increase the signal-to-noise (S/N) ratio, a beam-smoothing was applied, i.e., the map was smoothed with a Gaussian kernel of the size 7.′′ 5 (FWHM). Therefore, the angular resolution of the final image is 10.′′ 6 (0.02 pc at 450 pc). The gridding was done with a cell size of 1.′′ 5. The resulting 1σ rms noise level in the final co-added map is ∼ 0.06 Jy beam−1 . 2.2. Molecular-line observations The spectral-line observations of C17 O(2 − 1), DCO+ (4 − 3), N2 H+ (3 − 2), and N2 D+ (3 − 2) towards nine Orion B9 cores, and H13 CO+ (4 − 3) observations towards three cores, were carried out on 2–3, 8–9, and 11–12 September, and 13 and 18–19 November 2009 with APEX. The target cores with their physical properties derived in Paper II are listed in Table 1. In Table 1 we give the LABOCA submm peak positions of the cores after adjusting the pointing. The target positions of our molecular-line observations can be found in Table 1 of Paper II. The observed transitions, their spectroscopic properties, and observational parameters are listed in Table 2. The N2 H+ (3−2) observations were already presented in Paper II, but are re-analysed in the present paper. As frontend for the C17 O(2−1) and N2 D+ (3−2) observations, we used APEX-1 of the SHFI (Swedish Heterodyne Facility Instrument; Belitsky et al. 2007; Vassilev et al. 2008a,b). APEX-1 operates in single-sideband (SSB) mode using sideband separation mixers, and it has a sideband rejection ratio > 10 dB. The SHFI sidebands are separated by 12 GHz, i.e., ±6 GHz around the local oscillator (LO) frequency. Therefore, the centre frequency for the image band is 12 GHz above or below the observing frequency depending on whether the receiver tuning is optimised for operation in the upper or lower sideband (USB or LSB), respectively. For the H13 CO+ (4 − 3), DCO+ (4 − 3), and N2 H+ (3 − 2) observations the frontend used was APEX-2, which has similar characteristics as APEX-1. The backend for all observations was the Fast Fourier Transfrom Spectrometer (FFTS; Klein et al. 2006) with two 1 GHz units. Both units were divided into 8 192 channels. The two units were connected to one receiver each, thus providing 1 GHz bandwidth for two receivers simultaneously. Miettinen et al.: Dense cores in Orion B9 Fig. 1. Spitzer/MIPS 24-µm image of the central part of Orion B9 overlaid with contours showing the LABOCA 870-µm dust continuum emission. The contours go from 0.1 (∼ 3.3σ) to 1.0 Jy beam−1 in steps of 0.1 Jy beam−1 . The 24-µm image is shown with a logarithmic scaling to improve the contrast between bright and faint features. The colour bar indicates the 24-µm intensity scale in units of MJy sr−1 . The green plus signs show the target positions of our molecular-line observations (i.e., the submm peak positions of the LABOCA map before adjusting the pointing; see text). The green line shows the base and northwest tip of the HH 92 jet driven by IRAS 05399-0121 (Bally et al. 2002). The 0.2-pc scale bar and the effective LABOCA beam HPBW (∼ 20′′ ) are shown in the bottom left (Paper I). The observations were performed in the wobbler-switching mode with a 100′′ azimuthal throw (symmetric offsets) and a chopping rate of 0.5 Hz. The telescope pointing and focus corrections were checked by continuum scans on the planets Mars and Uranus, and the pointing was found to be accurate to ∼ 3′′ . Calibration was done by the chopper-wheel technique, and the output intensity scale given by the system is T A∗ , the antenna temperature corrected for atmospheric attenuation. The observed intensities were converted to the main-beam brightness temperature scale by T MB = T A∗ /ηMB , where ηMB = Beff /Feff is the main beam efficiency, and Beff and Feff are the beam and forward efficiencies, respectively. The absolute calibration uncertainty is estimated to be 10%. The spectra were reduced using the CLASS90 programme of the IRAM’s GILDAS software package3. The individual spectra were averaged and the resulting spectra were Hanning-smoothed in order to improve the S/N ratio of the data. A first- or thirdorder polynomial was applied to correct the baseline in the final spectra. The resulting 1σ rms noise levels are ∼ 20 − 80 mK at the smoothed resolutions. As shown in the second last column of Table 2, the on-source integration times were different between different sources. This is particularly the case for C17 O(2 − 1), 3 http://www.iram.fr/IRAMFR/GILDAS N2 D+ (3 − 2), and H13 CO+ (4 − 3) where tint varies by a factor of ∼ 3 − 7. This explains the variations in the rms noise values (up to a factor of 2.4 in the C17 O data). The J = 2 − 1 transition of C17 O contains nine hyperfine (hf) components. We fitted this hf structure using “method hfs” of CLASS90 to derive the LSR velocity (vLSR ) of the emission, and FWHM linewidth (∆v). The hf line fitting can also be used to derive the line optical thickness, τ. However, in all spectra the hf components are mostly blended together and thus the optical thickness could not be reliably determined. For the rest frequencies of the hf components, we used the values from Ladd et al. (1998; Table 6 therein). The adopted central frequency, 224 714.199 MHz, is that of the JF = 29/2 → 17/2 hf component which has a relative intensity of Ri = 31 . Also, rotational lines of H13 CO+ and DCO+ also have hf structure (e.g., Schmid-Burgk et al. 2004; Caselli & Dore 2005). The J = 4 − 3 transition of DCO+ is split up into six hf components. To fit this hf structure, we used the rest frequencies from the CDMS database (Müller et al. 2005). The adopted central frequency of DCO+ (4 − 3), 288 143.855 MHz, is that of the JF = 45 → 34 hf component which has a relative intensity of Ri = 11 27 . We note that this frequency is 2.7 kHz lower than the value determined by Caselli & Dore (2005; their Table 5). The frequency interval of the hf 3 Miettinen et al.: Dense cores in Orion B9 components for DCO+ (4 − 3) is very small. Therefore, the lines overlap significantly which causes the hf structure to be heavily blended. In general, for the Jupper ≥ 3 lines of DCO+ the hf components are so heavily blended that, even determined through a single Gaussian fit, the linewidth is not expected to be significantly overestimated (Caselli & Dore 2005). The hf structure of H13 CO+ lines is more complicated because both the 1 H and 13 C nuclei have a nuclear spin of I = 1/2, and the nuclear magnetic spin couples to rotation. To our knowledge, the rest frequencies of the H13 CO+ (4 − 3) hf components have not been published. We thus fitted these lines using a single Gaussian fit to derive the values of vLSR and ∆v. As in the case of DCO+ (Jupper ≥ 3), the ∆v thus determined is not expected to be significantly overestimated because of the strong blending of the hf components. The central frequency used was 346 998.344 MHz (CDMS), which is 3.0 kHz lower than the value determined by Schmid-Burgk et al. (2004; Table 3 therein). The J = 3 − 2 transitions of both N2 H+ and N2 D+ contain 38 hf components. The hf lines were fitted using the rest frequencies from Pagani et al. (2009b; Tables 4 and 10 therein). The adopted central frequencies of N2 H+ (3−2) and N2 D+ (3−2), 279 511.832 and 231 321.912 MHz, are those of the JF1 F = 345 → 234 hf 11 component which has a relative intensity of Ri = 63 . Also in these cases, the hf components are blended and thus the value of τ could not be reliably determined through hf fitting. 3. Observational results 3.1. SABOCA 350-µm emission The 350-µm SABOCA map is shown in Fig. 2. Almost all the cores detected with LABOCA show also clear 350 µm emission. The exceptions are SMM 5 and Ori B9 N for which the 350-µm peak flux densities are at the levels 3.5σ (210 mJy) and 3σ (180 mJy), respectively. With the peak flux density of 60.5σ (3630 mJy), SMM 3 is by far the strongest 350-µm source in the region. Four of the cores are resolved into at least two emission peaks in the SABOCA image. In the case of SMM 3, there is a 350-µm condensation, we call SMM 3b (4.2σ or 250 mJy), at about 36′′ east of the “main” source. At about 17′′ from SMM 3b, there is another 4σ (240 mJy) emission peak, designated here as SMM 3c. The 870 µm emission of SMM 3 extends to the direction of the subcondensations SMM 3b and 3c, and both of them lie either within or at the borderline of the 3.3σ (99 mJy) 870-µm contour. SMM 4 is resolved into two condensations. The western condensation coincides with the LABOCA peak position, whereas the eastern one, SMM 4b, is coincident with a Spitzer 24-µm source near SMM 4 (see Sect. 5.7.2). The elongated core SMM 6 is resolved into at least three subcondensations. The northwesternmost condensation corresponds to LABOCA peak of SMM 6. Finally, SMM 7 shows a hint of substructure in the western part of the core, where the 870 µm emission is above 3.3σ (99 mJy) level. SMM 3, 4, 6, 7, and their substructure will be discussed further in Sect. 5.7. There are also two 4 − 4.5σ (240 – 270 mJy) emission peaks at α2000.0 = 05h 42m 47.4s, δ2000.0 = −01◦ 17′ 15′′ (south of SMM 3b) and α2000.0 = 05h 43m 24.7s, δ2000.0 = −01◦14′ 35′′ (southeast of SMM 7), and a few 3.5 −4σ (210 – 240 mJy) peaks to the east of SMM 6 at α2000.0 = 05h 43m 16.5s, δ2000.0 ≃ −01◦ 19′ 00′′ . There is some weak LABOCA emission just slightly north of the latter peaks. However, because these “additional” sources are below 5σ (300 mJy), were not detected by LABOCA at 870 µm, and 4 were not identified by the SIMBAD Astronomical Database4 , they may well be unreal and are not discussed further in this paper. To identify and extract the cores from the SABOCA map, we employed the commonly used two-dimensional clumpfind algorithm, clfind2d, developed by Williams et al. (1994). The clfind2d routine determines the peak position, the FWHM size (not corrected for beam size), and the peak and total integrated flux density of the source based on specified contour levels. The algorithm requires two configuration parameters: i) the intensity threshold, i.e., the lowest contour level, T low , which determines the minimum emission to be included into the source; and ii) the contour level spacing, ∆T , which determines the required “contrast” between two sources to be considered as different objects. We set both parameters to 3σ (180 mJy). The selected 3σ contour levels turned out to give the best agreement with the identification by eye. With these parameter settings, the SMM 6c and 6d condensations are treated as a single source by clumpfind. The J2000.0 coordinates of√ the peak 350 µm emission, source effective radius (Reff = A/π, where A is the projected area within the 3σ contour), and peak and integrated flux densities are listed in Cols. (2)–(6) of Table 3. In Col. (7), we also list the flux densities measured in a 40′′ diameter aperture from the SABOCA map smoothed to the resolution of our LABOCA data. These flux densities were determined using the CRUSH-2 programme. The effective radius is only given for sources which are larger than the beam size. q The total flux density uncertainty was derived from σ(S λ ) = σ2cal + σ2S , where σcal is the absolute calibration error (adopted to be 30% of flux density), and σS is the uncertainty in the flux-density determination based on the rms noise near the source area. The 1σ rms noise in the smoothed SABOCA map, σ = 0.08 Jy beam−1 , is slightly higher than in the original map. We note that the data reduction produces negative artefacts (“holes”) around regions of bright emission, most notably around SMM 3. This decreases the source’s peak intensity, introducing an additional uncertainty in the flux density. The uncertainties due to negative bowls are neglected in the subsequent analysis. 3.2. Spectra The Hanning-smoothed spectra are shown in Fig. 3. The C17 O(2 − 1) line is clearly detected towards all sources except SMM 4 where only a low-velocity component is observed at vLSR ≃ 1.7 km s−1 (hereafter, SMM 4-LVC, where LVC stands for low-velocity component). We note that SMM 4 shows strong NH3 (1, 1) emission at about 9.1 km s−1 (Paper II). IRAS 054050117 (henceforth, IRAS05405 etc.) shows two additional velocity components at ∼ 1.3 and 3.0 km s−1 , and Ori B9 N shows an additional and wide C17 O(2 − 1) line at ∼ 1.9 km s−1 . There is also a hint of C17 O(2−1) emission at the systemic velocity of the region (∼ 9 km s−1 ) in the spectrum towards SMM 7, although the radial velocity of SMM 7, as determined from NH3 (1, 1) measurements in Paper II, is 3.6 km s−1 . The hf-structure of the C17 O(2 − 1) line is partially resolved in IRAS05399, SMM 1 and 3, and Ori B9 N. The H13 CO+ (4 − 3) observations were carried out only towards three sources (IRAS05399, SMM 1, and SMM 4). The line is only detected in IRAS05399 and SMM 4-LVC. The DCO+ (4 − 3) line is detected towards all the other sources except IRAS05405, SMM 4, and SMM 5. The line is also quite 4 http://simbad.u-strasbg.fr/simbad/ Miettinen et al.: Dense cores in Orion B9 Table 1. Source list. Source IRAS 05399-0121 SMM 1 SMM 3 IRAS 05405-0117 SMM 4 SMM 5 SMM 6 Ori B9 N SMM 7 α2000.0 a [h:m:s] 05 42 27.5 05 42 30.5 05 42 45.2 05 43 02.7 05 43 04.0 05 43 04.6 05 43 05.2 05 43 05.7 05 43 22.2 δ2000.0 a [◦ :′ :′′ ] -01 20 00 -01 20 55 -01 16 13 -01 16 31 -01 15 54 -01 17 17 -01 18 48 -01 14 51 -01 13 56 T kin [K] 13.5 ± 1.6 11.9 ± 0.9 11.3 ± 0.8 11.3 ± 0.6 13.9 ± 0.8 11.3 ± 0.7 11.0 ± 0.4 13.4 ± 1.3 9.4 ± 1.1 M [M⊙ ] 7.8 ± 0.5 11.1 ± 0.3 7.8 ± 0.6 2.8 ± 0.4 2.8 ± 0.3 1.9 ± 0.4 8.2 ± 1.1 2.3 ± 0.4 3.6 ± 1.0 N(H2 )b [1022 cm−2 ] 4.2 ± 0.9/2.8 ± 0.6 2.8 ± 0.4/2.7 ± 0.4 8.4 ± 1.1/2.5 ± 0.3 1.4 ± 0.1/1.2 ± 0.1 1.4 ± 0.1/1.1 ± 0.1 1.2 ± 0.1/0.9 ± 0.1 2.0 ± 0.1/1.7 ± 0.1 0.9 ± 0.1/0.9 ± 0.1 3.4 ± 0.9/2.1 ± 0.5 hn(H2 )i [104 cm−3 ] 5.5 ± 1.3 5.3 ± 0.9 10.5 ± 2.1 3.8 ± 0.5 3.8 ± 0.4 2.5 ± 0.5 2.2 ± 0.3 2.1 ± 0.4 4.8 ± 1.3 Class 0/I prestellar 0 0 0 prestellar ? prestellar prestellar ? prestellar Notes. Columns (2) and (3) give the equatorial coordinates [(α, δ)2000.0 ]. Columns (4)–(7) list the gas kinetic temperature, core mass, beamaveraged peak H2 column density, and the volume-averaged H2 number density, respectively. In the last column we give the comments on the source classification. The virial parameter of the starless cores SMM 5 and Ori B9 N is αvir & 2, and thus it is unclear if they are prestellar (Paper II). (a) These coordinates refer to the LABOCA peak positions after adjusting the pointing by (∆α, ∆δ) = (+1.′′ 72, −9.′′ 97). The coordinates of our molecular-line observation target positions can be found in Table 1 of Paper II. (b) The first value is calculated towards the revised LABOCA peak position by using the temperature derived towards the line observation target position. The latter N(H2 ) value refers to the position used in molecular-line observations. Table 2. Observed spectral-line transitions and observational parameters. Transition C17 O(2 − 1) N2 D+ (3 − 2) N2 H+ (3 − 2) DCO+ (4 − 3) H13 CO+ (4 − 3) ν [MHz] 224 714.199b 231 321.912c 279 511.832c 288 143.855d 346 998.344e Eu /kB [K] 16.2 22.2 26.8 34.6 41.6 ncrit [cm−3 ] 9.0 × 103 1.7 × 106 2.9 × 106 4.7 × 106 8.2 × 106 HPBW [′′ ] 27.8 27.0 22.3 21.7 18.0 ηMB 0.75 0.75 0.74 0.74 0.73 T sys [K] 255–288 235–335 211–359 170–172 362–521 PWV [mm] 0.8–1.4 0.2–3.8 0.04–1.4 0.01–0.5 0.9–1.3 Channel spacinga [kHz] [km s−1 ] 122.07 0.16 122.07 0.16 122.07 0.13 122.07 0.13 122.07 0.11 tint [min] 2.9–19.5 11.5–37.5 2.7–5.5 4.1–6.5 8.5–22 rms [mK] 24–58 22–35 32–84 23–29 25–55 Notes. Columns (2)–(4) give the rest frequencies of the observed transitions (ν), their upper-state energies (Eu /kB , where kB is the Boltzmann constant), and critical densities. Critical densities were calculated at T = 10 K using the collisional-rate data available in the Leiden Atomic and Molecular Database (LAMDA; http://www.strw.leidenuniv.nl/∼moldata/) (Schöier et al. 2005). For N2 D+ , we used the Einstein A−coefficient from Pagani et al. (2009b) and the same collisional rate as for N2 H+ . Columns (5)–(12) give the APEX beamsize (HPBW) and the main beam efficiency (ηMB ) at the observed frequencies, the SSB system temperatures during the observations (T sys in T MB scale, see text), the amount of PWV, channel widths (both in kHz and km s−1 ) of the original data, the on-source integration times per position (tint ), and the 1σ rms noise at the smoothed resolution. (a) The original channel spacings. The final spectra were Hanning-smoothed which divides the number of channels by two.(b) From Ladd et al. (1998). (c) From Pagani et al. (2009b).(d) From the Jet Propulsion Laboratory (JPL; http://spec.jpl.nasa.gov/) spectroscopic database (Pickett et al. 1998).(e) From the Cologne Database for Molecular Spectroscopy (CDMS; http://www.astro.uni-koeln.de/cdms/catalog) (Müller et al. 2005). weak (4σ) in Ori B9 N. Again, clear emission from an additional velocity component can be seen towards SMM 4 and Ori B9 N. As already presented in Paper II, N2 H+ (3 − 2) emission is clearly seen in all sources with the exception of target position near Ori B9 N where the line is very weak. It should be noted, however, that towards SMM 4 only the LVC is detected. Ori B9 N shows an additional velocity component at about 1.8 km s−1 . Finally, the N2 D+ (3 − 2) line was detected in IRAS05399, SMM 1, SMM 3, IRAS05405, SMM 4-LVC, and SMM 6. In addition, the line appears at ∼ 3 km s−1 towards IRAS05405. 3.2.1. Other line detections All the observed sources except SMM 5 show additional spectral lines in the frequency band covering the N2 D+ (3 − 2) transition. The line identification was done by using Weeds, which is an extension of CLASS (Maret et al. 2011), and the JPL and CDMS spectroscopic databases. We used the LTE modelling application of Weeds to check if all predicted lines of a candidate molecule are present in the observed spectrum. In some cases, we were able to reject some line candidates on the basis of non-detection of other transitions expected at nearby frequencies. The JKa ,Kc = 40,4 − 30,3 transition of ortho-D2 CO at ∼ 231.4 GHz was detected towards IRAS05399 and SMM 1 in the LSB (see Fig. 4). Also, the J = 2 − 1 transition of C18 O at ∼ 219.56 GHz (LSB) was detected in the image sideband towards all sources except SMM 5 [marked with “(i)”; Fig. 5]. An additional velocity component is detected towards IRAS05405 and SMM 4. Moreover, in the spectrum towards Ori B9 N several distinct velocity components are detected. The possible fourth velocity component of C18 O could, in principle, be blended with the line at ∼ 231.06 GHz, which can be assigned to either OCS(J = 19 − 18) (231 060.9830 MHz, Eu /kB = 110.9 K) or CH3 NH2 -E(JKa ,Kc = 72,5 − 71,5 ) (231 060.6041 MHz, Eu /kB = 75.6 K) because these two transitions are also blended. However, these can almost certainly be excluded because of the high transition energies involved, and because both species are expected to be formed by grain-surface chemistry. Detection of these species in a cold core is therefore very unlikely. Note that the lines “leaking” from the rejected image band are heavily attenuated by the sideband filter. Therefore, we cannot establish the correct intensity scale for the C18 O(2 − 1) lines. 5 Miettinen et al.: Dense cores in Orion B9 Fig. 2. SABOCA 350-µm image (smoothed to 10.′′ 6 resolution) of the central part of Orion B9 (colour scale and green contours). The image is shown with a square root scaling, and the colour bar indicates the flux density in units of Jy beam−1 . The rms level is 0.06 Jy beam−1 (1σ). The first 350-µm contour and the separation between contours is 3σ. The white contours show the LABOCA 870-µm dust continuum emission as in Fig. 1. The yellow plus signs show the target positions of our molecular-line observations. The 0.2-pc scale bar and beam HPBW are shown in the bottom right. Table 3. The 350-µm core properties. Name SMM 3 SMM 3b SMM 3c IRAS 05405-0117 SMM 4 SMM 4b SMM 5 SMM 6 SMM 6b SMM 6c SMM 6d Ori B9 N SMM 7 SMM 7b Peak position α2000.0 [h:m:s] δ2000.0 [◦ :′ :′′ ] 05 42 45.3 -01 16 16 05 42 47.6 -01 16 24 05 42 48.6 -01 16 32 05 43 03.0 -01 16 31 05 43 04.0 -01 15 52 05 43 05.7 -01 15 57 05 43 03.8 -01 16 59 05 43 04.9 -01 18 49 05 43 06.7 -01 19 02 05 43 07.9 -01 19 13 05 43 08.6 -01 19 22 05 43 05.6 -01 14 55 05 43 22.7 -01 13 55 05 43 21.3 -01 14 02 Reff [′′ ] 13.3 ... ... 7.6 5.8 7.5 ... 8.2 7.0 6.2b ... ... 10.1 ... peak S 350 [Jy beam−1 ] 3.63 0.25 0.24 0.47 0.25 0.49 0.21 0.33 0.27 0.21 0.23 0.18 0.32 0.21 S 350 [Jy] 5.4 ± 1.6 ... ... 0.5 ± 0.2 0.2 ± 0.1 0.5 ± 0.2 ... 0.5 ± 0.2 0.3 ± 0.1 0.2 ± 0.1b ... ... 0.7 ± 0.2 ... ′′ 40 a S 350 [Jy] 3.9 ± 1.2 ... ... 0.6 ± 0.2 0.7 ± 0.3 ... 0.5 ± 0.2 0.7 ± 0.3 ... ... ... 0.5 ± 0.2 0.8 ± 0.3 ... Notes. (a) Flux density measured in a 40′′ diameter aperture from the SABOCA map smoothed to the resolution of the LABOCA map. (b) These values refer to the combined size and flux density of SMM 6c and 6d. In Table 4 we list the detected extra transitions, their rest frequencies, and upper-state energies. The rest frequencies and upper-state energies were assigned using JPL, CDMS, and Splatalogue5 (Remijan et al. 2007) spectroscopic databases. The D2 CO detections are discussed further in Sect. 5.6. 5 6 http://www.splatalogue.net/ 3.3. Spectral-line parameters The spectral line parameters are given in Table 5. In this table we give the radial velocity (vLSR ), FWHM linewidth (∆v), peak inR tensity (T MB ), integrated intensity ( T MB dv), peak optical thickness of the line (τ0 ), and excitation temperature (T ex ). For nondetections, the 3σ upper limit on the line intensity is given. The values of vLSR and ∆v for C17 O, DCO+ , N2 H+ , and N2 D+ were Miettinen et al.: Dense cores in Orion B9 Fig. 3. Smoothed C17 O(2 − 1), H13 CO+ (4 − 3), DCO+ (4 − 3), N2 H+ (3 − 2), and N2 D+ (3 − 2) spectra. Overlaid on the C17 O, DCO+ , N2 H+ , and N2 D+ spectra are the hf-structure fits. The relative velocities of individual hf components in these transitions are labelled with a short bar on the spectra towards IRAS05399. The H13 CO+ spectra are overlaid with a single Gaussian fit. The fits to the lines at the systemic velocity ∼ 9 km s−1 are shown as green lines, whereas the red lines show fits to the lines at lower velocities. Note that H13 CO+ (4 − 3) observations were carried out only towards three sources and the line is detected only towards IRAS05399 and SMM 4 (at ∼ 1.5 km s−1 ). Table 4. Other detected species/transitions. Species/ transitiona C18 O(J = 2 − 1)b ortho-D2 CO(JKa ,Kc = 40,4 − 30,3 ) ν [MHz] 219 560.3600 (JPL) 231 410.234 (CDMS) Eu /kB [K] 15.8 27.9 Notes. (a) For asymmetric top molecules such as D2 CO, Ka and Kc refer to the projection of the angular momentum along the a and c principal axes. The detected o-D2 CO transition exbits a-type selection rules (∆Ka = 0, ±2, . . . and ∆Kc = ±1, ±3, . . .) (Gordy & Cook 1984). (b) Seen in the image band. derived through fitting the hf structure. For the other lines these parameters were derived by fitting a single Gaussian to the line R profile. Also, the values of T MB and T MB dv were determined from Gaussian fits. The integrated C17 O(2 − 1) line intensities of IRAS05399, SMM 1, 3, and 7, and Ori B9 N were calculated over the velocity range given in Col. (6) of Table 5 because the hf structure of the line is partially resolved in these cases. The uncertainties reported in vLSR and ∆v represent the formal 1σ errors R determined by the fitting routine, whereas those in T MB and T MB dv also include the 10% calibration uncertainty. We used RADEX6 (van der Tak et al. 2007) to determine the values of τ0 and T ex for the lines of C17 O, H13 CO+ , DCO+ , and N2 H+ . 6 http://www.strw.leidenuniv.nl/∼moldata/radex.html 7 Miettinen et al.: Dense cores in Orion B9 Fig. 3. continued. RADEX modelling is described in more detail in Sect. 4.4 where we discuss the determination of molecular column densities. For the rest of the lines, τ0 was determined through Weeds modelling (Sect. 4.4), and the associated error was estimated from the 10% calibration uncertainty. The lines appear to be optically thin in most cases, except that the N2 H+ lines are mostly optically thick (τ0 > 1). The T ex values for C17 O(2−1) are close to T kin , indicating that the lines are nearly thermalised. The T ex [H13 CO+ (4−3)] values are found to be similar to T ex [DCO+ (4 − 3)]. We note that the T ex values obtained for the J = 3 − 2 transition of N2 H+ (3.8–5.0 K) are mostly comparable to the values T ex = 4.6 − 5.5 K determined or assumed in Paper II. For N2 D+ (3 − 2) we adopt as T ex the value derived for N2 H+ (3 − 2). 8 4. Analysis and results 4.1. Core properties derived from 350 µm emission We used the 350-µm dust continuum data to determine the mass, peak beam-averaged column density of H2 , and volumeaveraged H2 number density of the cores and their condensations. The formulas for the mass and H2 column density can be found in Paper I [Eqs. (2) and (3) therein]. Note that in Eq. (3) peak of Paper I, the peak surface brightness is Iλdust = S λ /Ωbeam , where Ωbeam is the solid angle of the telescope beam. As the dust temperature of the sources, we used the gas kinetic temperatures listed in Col. (4) of Table 1, and assumed that T dust = T kin . For the subcondensations, such as SMM 6b, 6c, etc., it was assumed that T dust is the same as in the “main” core. This assumption may not be valid, however, because in most cases the subcondensations lie outside the 40′′ beam of the NH3 measurements used to derive T kin in the main cores. Moreover, the temperature of an individual small condensation may be lower than the temperature of the parent core because of more effective shield- Miettinen et al.: Dense cores in Orion B9 Fig. 3. continued. Fig. 5. C18 O(2 − 1) lines at 219 560.3568 MHz (JPL) arising from the image sideband. The x-axis range is different in the spectrum towards Ori B9 N, and the lines are overlaid with Gaussian fits for a better illustration. The blended OCS and CH3 NH2 lines at ∼ 231.06 GHz marked on the spectrum towards Ori B9 N can be excluded on the basis of high upper-state energies and formation chemistry of the species (see text). The line is likely to be caused by an additional C18 O(2 − 1) velocity component. ing from the external radiation field. If the temperature is lower than assumed the mass and column density will be underestimated. As the dust opacity per unit dust mass at 350 µm, κ350 µm , we used the value 1.0 m2 kg−1 which was extrapolated from the Ossenkopf & Henning (1994, hereafter OH94) model describing graphite-silicate dust grains that have coagulated and accreted thick ice mantles over a period of 105 yr at a gas den- sity of nH = n(H) + 2n(H2 ) ≃ 2n(H2 ) = 105 cm−3 . The same dust model was adopted in Paper I (κ870 µm ≃ 0.17 m2 kg−1 ), and is expected to be a reasonable model for cold, dense molecular cloud cores7 . We note that the sub-mm opacities recently 7 For comparison, for dust grains covered by thin ice mantles at a density 105 cm−3 , κ350 µm ≃ 0.78 m2 kg−1 . The κ350 µm value we have adopted is the same as in the OH94 model of grains with thin ice man9 Miettinen et al.: Dense cores in Orion B9 Table 5. Spectral-line parameters. Source Transition IRAS 05399-0121 C17 O(2 − 1) H13 CO+ (4 − 3) DCO+ (4 − 3) N2 H+ (3 − 2) N2 D+ (3 − 2) o-D2 CO(40,4 − 30,3 ) C17 O(2 − 1) H13 CO+ (4 − 3) DCO+ (4 − 3) N2 H+ (3 − 2) N2 D+ (3 − 2) o-D2 CO(40,4 − 30,3 ) C17 O(2 − 1) DCO+ (4 − 3) N2 H+ (3 − 2) N2 D+ (3 − 2) C17 O(2 − 1) C17 O(2 − 1) C17 O(2 − 1) DCO+ (4 − 3) N2 H+ (3 − 2) N2 D+ (3 − 2) N2 D+ (3 − 2) C17 O(2 − 1) H13 CO+ (4 − 3) DCO+ (4 − 3) N2 H+ (3 − 2) N2 D+ (3 − 2) C17 O(2 − 1) DCO+ (4 − 3) N2 H+ (3 − 2) N2 D+ (3 − 2) C17 O(2 − 1) DCO+ (4 − 3) N2 H+ (3 − 2) N2 D+ (3 − 2) C17 O(2 − 1) C17 O(2 − 1) DCO+ (4 − 3) DCO+ (4 − 3) N2 H+ (3 − 2) N2 H+ (3 − 2) N2 D+ (3 − 2) C17 O(2 − 1) C17 O(2 − 1) DCO+ (4 − 3) N2 H+ (3 − 2) N2 D+ (3 − 2) SMM 1 SMM 3 IRAS 05405-0117 2nd v-comp. 3rd v-comp. 3rd v-comp. SMM 4 (2nd v-comp.) 2nd v-comp. 2nd v-comp. 2nd v-comp. 2nd v-comp. SMM 5 SMM 6 Ori B9 N 2nd v-comp. 2nd v-comp. 2nd v-comp. SMM 7 “9 km s−1 ”-comp. vLSR [km s−1 ] 8.68 ± 0.02 8.62 ± 0.08 8.58 ± 0.03 8.73 ± 0.02 8.46 ± 0.04 8.70 ± 0.22 9.20 ± 0.01 ... 9.31 ± 0.01 9.31 ± 0.02 9.20 ± 0.03 9.21 ± 0.06 8.68 ± 0.06 8.54 ± 0.03 8.57 ± 0.03 8.39 ± 0.04 9.20 ± 0.04 1.30 ± 0.33 2.98 ± 0.33 ... 9.30 ± 0.03 8.93 ± 0.15 2.98 ± 0.20 1.67 ± 0.05 1.53 ± 0.08 1.65 ± 0.03 1.62 ± 0.78 1.49 ± 0.04 9.37 ± 0.03 ... 9.43 ± 0.07 ... 9.51 ± 0.04 9.47 ± 0.02 9.52 ± 0.03 9.25 ± 0.03 9.29 ± 0.33 1.85 ± 0.33 9.42 ± 0.19 2.11 ± 0.11 8.80 ± 0.16 1.81 ± 0.11 ... 3.75 ± 0.02 9.04 ± 0.07 3.70 ± 0.04 4.01 ± 0.14 ... ∆v [km s−1 ] 0.74 ± 0.06 0.82 ± 0.16 0.69 ± 0.06 0.71 ± 0.07 0.53 ± 0.02 2.20 ± 0.46 0.93 ± 0.03 ... 0.47 ± 0.05 0.66 ± 0.09 0.75 ± 0.07 0.67 ± 0.16 0.58 ± 0.09 0.42 ± 0.16 0.85 ± 0.09 0.53 ± 0.11 0.54 ± 0.04 0.58 ± 1.08 0.54 ± 1.08 ... 0.69 ± 0.06 0.53 ± 0.34 0.53 ± 0.56 0.92 ± 0.19 0.43 ± 0.15 0.49 ± 0.09 0.61 ± 21.10 0.54 ± 0.10 0.54 ± 0.01 ... 0.44 ± 0.17 ... 0.54 ± 0.04 0.42 ± 0.04 0.65 ± 0.09 0.56 ± 0.10 0.54 ± 1.08 1.39 ± 1.08 0.47 ± 1.38 0.78 ± 0.31 0.44 ± 0.24 0.85 ± 0.11 ... 1.04 ± 0.05 0.55 ± 1.74 0.42 ± 0.10 0.67 ± 0.27 ... T MB [K] 0.79 ± 0.18 0.09 ± 0.02 0.33 ± 0.04 0.88 ± 0.11 0.28 ± 0.03 0.08 ± 0.03 0.96 ± 0.17 < 0.13 0.55 ± 0.07 0.65 ± 0.07 0.38 ± 0.04 0.15 ± 0.02 0.35 ± 0.05 0.20 ± 0.02 0.62 ± 0.09 0.21 ± 0.03 0.30 ± 0.03 0.17 ± 0.02 0.20 ± 0.02 < 0.08 0.66 ± 0.09 0.06 ± 0.01 0.05 ± 0.01 0.32 ± 0.06 0.15 ± 0.02 0.21 ± 0.04 0.32 ± 0.07 0.24 ± 0.03 0.22 ± 0.02 < 0.07 0.16 ± 0.03 < 0.11 0.28 ± 0.03 0.37 ± 0.05 0.51 ± 0.06 0.33 ± 0.03 0.30 ± 0.05 0.29 ± 0.05 0.04 ± 0.01 0.09 ± 0.01 0.05 ± 0.02 0.09 ± 0.02 < 0.07 0.66 ± 0.09 0.10 ± 0.06 0.22 ± 0.02 0.14 ± 0.02 < 0.07 R T MB dva [K km s−1 ] 1.27 ± 0.14 [6.40, 9.64] 0.08 ± 0.02 0.25 ± 0.03 1.00 ± 0.11 [92.6%] 0.14 ± 0.02 [91.6%] 0.18 ± 0.04 1.85 ± 0.20 [7.00, 11.33] ... 0.42 ± 0.05 0.70 ± 0.08 [92.6%] 0.40 ± 0.04 [92.6%] 0.10 ± 0.02 0.51 ± 0.09 [6.56, 9.52] 0.09 ± 0.01 0.67 ± 0.08 [92.6%] 0.17 ± 0.03 [91.6%] 0.17 ± 0.03 [78.7%] 0.14 ± 0.02 [78.7%] 0.13 ± 0.02 [78.7%] ... 0.56 ± 0.07 [90.9%] 0.04 ± 0.03 [91.5%] 0.05 ± 0.02 [92.4%] 0.61 ± 0.08 0.07 ± 0.02 0.19 ± 0.03 0.27 ± 0.07 [92.6%] 0.17 ± 0.03 [90.9%] 0.14 ± 0.02 [78.7%] ... 0.14 ± 0.02 [92.6%] ... 0.15 ± 0.03 [69.4%] 0.19 ± 0.02 0.44 ± 0.06 [92.6%] 0.26 ± 0.03 [90.9%] 0.20 ± 0.03 [96.0%] 0.70 ± 0.08 0.02 ± 0.02 0.09 ± 0.02 0.05 ± 0.02 [92.6%] 0.13 ± 0.03 [92.6%] ... 0.82 ± 0.09 [1.38, 4.95] 0.07 ± 0.03 [69.4%] 0.10 ± 0.02 0.17 ± 0.04 [92.6%] ... τ0 b,c 0.10 ± 0.05 0.11 ± 0.05 0.51 ± 0.10 2.80 ± 0.20 0.68 ± 0.10 0.01 ± 0.001 0.15 ± 0.05 ... 1.89 ± 0.20 2.43 ± 0.10 1.71 ± 0.20 0.02 ± 0.002 0.05 ± 0.03 0.24 ± 0.03 1.23 ± 0.10 0.59 ± 0.10 0.05 ± 0.03 ... ... ... 4.26 ± 0.10 0.16 ± 0.02 ... 0.06 ± 0.03 1.02 ± 0.60 0.76 ± 0.50 1.59 ± 0.50 1.50 ± 0.20 0.04 ± 0.03 ... 0.67 ± 0.04 ... 0.05 ± 0.03 3.47 ± 0.20 5.75 ± 0.35 2.92 ± 0.30 0.04 ± 0.03 0.04 ± 0.03 0.05 ± 0.02 0.14 ± 0.08 0.12 ± 0.03 0.23 ± 0.07 ... 0.15 ± 0.05 ... 0.94 ± 0.20 0.49 ± 0.07 ... T ex c [K] 12.8 ± 1.5 5.5 ± 0.5 4.9 ± 0.3 5.0 ± 0.2 5.0 ± 0.2d 18.6 11.3 ± 1.0 ... 4.5 ± 0.3 4.6 ± 0.3 4.6 ± 0.3d 18.6 11.0 ± 1.0 5.1 ± 0.3 4.9 ± 0.3 4.9 ± 0.3d 10.5 ± 0.5 ... ... ... 4.5 ± 0.3 4.5 ± 0.3d ... 9.7 ± 1.2 4.0 ± 0.2 4.0 ± 0.3 4.0 ± 0.3 4.0 ± 0.3d 10.1 ± 0.5 ... 3.8 ± 0.1 ... 9.7 ± 0.5 4.0 ± 0.3 4.2 ± 0.2 4.2 ± 0.2d 11.6 ± 1.0 11.7 ± 2.0 4.9 ± 0.3 4.6 ± 0.5 4.1 ± 0.2 4.1 ± 0.2 ... 8.9 ± 1.0 ... 3.9 ± 0.3 3.9 ± 0.2 ... Notes. (a) Integrated intensity is derived from a Gaussian fit or, in the case of some C17 O lines, by integrating over the velocity range indicated in brackets. The percentage in brackets indicates the contribution of hf component’s intensity lying within the Gaussian fit.(b) For C17 O, H13 CO+ , DCO+ , N2 H+ , and N2 D+ τ0 is the optical thickness in the centre of a hypothetical unsplit line (see text).(c) For C17 O, H13 CO+ , DCO+ , and N2 H+ the values of τ0 and T ex were estimated using RADEX with the kinetic temperature (T kin ) and the H2 density (hn(H2 )i). The value of τ0 for N2 D+ and o-D2 CO was estimated using CLASS/Weeds; see Sect. 4.4 for details.(d) T ex is assumed to be the same as for N2 H+ (3 − 2). calculated by Ormel et al. (2011) are comparable to the OH94 values (after 105 yr of coagulation). However, the dust opacities are likely to be uncertain by a factor of & 2 (e.g., OH94; Motte & André 2001; Ormel et al. 2011). For the average dustto-gas mass ratio, Rd ≡ hMdust /Mgas i, we adopted the canonical value 1/100. Finally, we assumed a He/H abundance ratio of 0.1, which leads to the mean molecular weight per H2 molecule of µH2 = 2.8. The integrated flux densities used to calculate the masses refer to core areas, A. Thus, in order to properly calculate the volume-average H2√ number density, hn(H2 )i, we use the effective radius, Reff = A/π, in the formula hn(H2 )i = hρi , µH2 mH (1) tles at a density 106 cm−3 . At 106 cm−3 with thick ice mantles κ350 µm rises to about 1.1 m2 kg−1 . 10   where hρi = M/ 4π/3 × R3eff is the mass density, and mH is the mass of a hydrogen atom. The results of the above calculations are presented in Table 6. The uncertainties in the derived parameters were propagated from the uncertainties in T kin . We note that the 1σ rms noise on our SABOCA map, ∼ 0.06 Jy beam−1 , corresponds to a 3σ mass detection limit of ∼ 0.1 M⊙ assuming T dust = 10 K. In terms of column density, the 3σ detection limit is about 2.8 × 1021 cm−2 . 4.2. Dust properties determined from 350 and 870 µm data We can use our observations at two submm wavelengths, 350 and 870 µm, to estimate the dust colour temperature, T dust , and dust emissivity spectral index, β [κλ = κ0 (λ0 /λ)β ]. Note that according to the Wien displacement law, the wavelength of the peak of the blackbody radiation curve is about 290 µm at 10 K. Miettinen et al.: Dense cores in Orion B9 Table 7. Dust colour temperature and emissivity index derived from 350-to-870 µm flux density ratio. Source SMM 3 IRAS 05405-0117 SMM 4 SMM 5 SMM 6 Ori B9 N SMM 7 ′′ ′′ 40 40 S 350 /S 870 2.4 ± 1.1 1.5 ± 0.6 1.4 ± 0.8 1.7 ± 0.9 1.2 ± 0.6 1.7 ± 0.9 1.6 ± 0.9 T dust a [K] +5.7 10.8−2.6 +2.6 8.7−1.7 +4.5 8.4−2.0 +4.3 9.0−2.2 +3.0 7.9−1.8 +4.3 9.0−2.2 +4.8 8.9−2.1 βb 1.8 ± 0.6 1.1 ± 0.6 0.5 ± 0.8 1.2 ± 0.8 0.8 ± 0.9 0.8 ± 0.8 1.7 ± 0.8 Notes. (a) Derived by assuming β = 1.9 (see text).(b) Calculated by assuming T dust = T kin as derived from NH3 observations. 4.3. SEDs Fig. 4. ortho-D2CO(40,4 − 30,3 ) spectra towards IRAS05399 and SMM 1. Table 6. The effective radius, mass, H2 column density, and volume-averaged H2 density derived from the 350 µm emission. Source SMM 3 SMM 3b SMM 3c IRAS 05405-0117 SMM 4 SMM 4b SMM 5 SMM 6 SMM 6b SMM 6c SMM 6d Ori B9 N SMM 7 SMM 7b Reff [pc] 0.03 ... ... 0.02 0.01 0.02 ... 0.02 0.02 0.01b ... ... 0.02 ... M [M⊙ ] 2.1 ± 0.8 ... ... 0.2 ± 0.1 0.04 ± 0.02 0.1 ± 0.04a ... 0.2 ± 0.1 0.1 ± 0.05a 0.1 ± 0.04a,b ... ... 0.6 ± 0.3 ... N(H2 ) [1022 cm−2 ] 10.4 ± 2.7 0.7 ± 0.2a 0.7 ± 0.2a 1.3 ± 0.3 0.4 ± 0.1 0.7 ± 0.1a 0.6 ± 0.1 1.0 ± 0.1 0.9 ± 0.1a 0.7 ± 0.1a 0.7 ± 0.1a 0.3 ± 0.1 1.9 ± 1.0 1.3 ± 0.7a hn(H2 )i [105 cm−3 ] 4.0 ± 1.5 ... ... 2.0 ± 1.0 0.9 ± 0.5 1.0 ± 0.4 ... 1.6 ± 0.8 1.3 ± 0.6 1.9 ± 0.7b ... ... 2.6 ± 1.3 ... Notes. (a) Calculated by assuming T dust is the same as for the “main” core.(b) These values include contributions from both SMM 6b and 6c. Therefore, the Rayleigh-Jeans (R-J) approximation, hν ≪ kB T , is not valid for our 350 and 870 µm data. In the following analysis, it is assumed that T dust and β are constant across the source. We first smoothed the SABOCA map to the resolution of our LABOCA data, and then calculated the flux densities at both wavelengths in a fixed 40′′ diameter aperture [Col. (7) of Table 3]. The resulting flux density ratios, ′′ ′′ 40 40 S 350 /S 870 , are given in Col. (2) of Table 7. A dust colour temperature can be determined by fixing the value of β [see, e.g., Eq. (3) of Shetty et al. (2009)]. The value of β in the OH94 thickice dust model we adopted earlier is about 1.9 over the wavelength range λ ∈ [250, 1300 µm] (see also Shirley et al. 2005). By adopting the value β = 1.9, we derive the T dust values in the range ∼ 7.9 − 10.8 K [see Col. (3) of Table 7]. To calculate β from the ratio of two flux densities at different wavelengths, an estimate for the dust temperature is needed. We assumed that T dust = T kin , and applied Eq. (5) of Shetty et al. (2009). The resulting values, β ≃ 0.5 − 1.8, are listed in Col. (4) of Table 7. We note that T kin measurements were obtained with a 40′′ resolution, whereas the above flux density ratio was determined at about 20′′ resolution. However, flux densities were measured in a 40′′ aperture, which matches the resolution of our NH3 data. With the aid of our new 350 µm data, we were able to refine some of the protostellar SEDs presented in Paper I. As in Paper I, the SEDs constructed from the 24, 70, 350, and 870-µm flux densities were fitted by the sum of two modified blackbody curves of cold and warm temperatures (the original version of the fitting routine was written by J. Steinacker). Again, the dust model used to fit the SEDs is the OH94 model of thick ice mantles as adopted earlier in Sect. 4.1, incorporating a dust-to-gas mass ratio of 1/100. The SEDs of SMM 3, IRAS05405, and SMM 4 are shown in Fig. 6. Note that in Paper I we utilised the IRAS flux densities to build the SED of IRAS05405. In the present paper we have ignored the IRAS data from the fit because the ∼arcminute resolution of the IRAS observations also includes emission from the nearby core SMM 4, confusing the emission from IRAS05405. This probably explains why the 100-µm flux density of IRAS05405 appears so high (∼ 19.7 Jy). The SED fitting results are shown in Table 8. Column (2) of Table 8 gives the mass of the cold component, which represents the mass of the cold envelope, Mcold ≡ Menv . In Cols. (3) – (5), we list the luminosities of the cold and warm component, and the bolometric luminosity, Lbol = Lcold + Lwarm . The temperature of the cold component, T cold , is given in Col. (6). Columns (7) and (8) give the Lcold /Lbol and Lsubmm /Lbol ratios, where the submm luminosity, Lsubmm , is defined to be the luminosity longward of 350 µm. In the last column of Table 8 we give the normalised envelope mass, Menv /L0.6 bol (Bontemps et al. 1996). The latter parameter, which is related to the outflow activity, decreases with time and can be used to further constrain the core evolutionary stage. We do not report the temperature and mass of the warm component, because i) the OH94 dust model of grains covered by thick ice mantles we have adopted may not be appropriate for the warm dust component; and ii) dust emission may not be optically thin at 24 and 70 µm, so the masses and temperatures of the warm component are not well constrained. In contrast, the luminosity of the warm component is not affected by these opacity effects. 4.4. Molecular column densities and fractional abundances The beam-averaged column densities of C17 O, H13 CO+ , DCO+ , and N2 H+ were derived using a one-dimensional spherically symmetric non-LTE radiative transfer code called RADEX (see Sect. 3.3). RADEX uses the method of mean escape probability for an isothermal and homogeneous medium. The molecular data files (collisional rates) used in the RADEX excitation analysis were taken from the LAMDA database (Schöier et al. 2005). The C17 O, H13 CO+ , DCO+ , and N2 H+ transitions are treated as 11 Miettinen et al.: Dense cores in Orion B9 Fig. 6. Spectral energy distributions of three protostellar cores in Orion B9 built from Spitzer 24 and 70 µm, SABOCA 350 µm, and LABOCA 870-µm flux densities. Open circles in the middle panel represent IRAS data points at 12, 25, 60, and 100 µm (not used in the fit). The solid lines correspond to the two-temperature model fits to the data. Error bars (1σ) are indicated for all data points, but are mostly smaller than the symbol size. Note the appearance of a 10-µm silicate absorption feature in the SED of IRAS05405, and the absorption “knee” at ∼ 30 µm (most notably towards IRAS05405) where a considerable change of κν occurs at the dust model used (OH94 and references therein). Table 8. Fitting results of the SEDs. Source SMM 3 IRAS 05405-0117 SMM 4 Notes. (a) Mcold a [M⊙ ] 8.2 ± 2.6 1.6 ± 0.5 2.0 ± 0.8 Lcold [L⊙ ] 0.3 ± 0.1 0.06 ± 0.02 0.07 ± 0.03 Lwarm [L⊙ ] 0.9 ± 0.1 2.3 ± 0.2 0.5 ± 0.02 12 T cold [K] 8.0 8.0 8.0 Lcold /Lbol Lsubmm /Lbol 0.3 ± 0.1 0.03 ± 0.01 0.1 ± 0.05 0.1 0.01 0.05 Menv /L0.6 bol [M⊙ /L0.6 ⊙] 7.4 ± 3.2 1.0 ± 0.3 2.7 ± 1.1 Mcold ≡ Menv .(b) Lbol = Lwarm + Lcold . a hypothetical unsplit transition. The input parameters in the offline mode of RADEX are the gas kinetic temperature, H2 number density, and the width (FWHM) and intensity of the spectral line. We used the values of T kin and hn(H2 )i listed in Table 1. However, we multiplied the densities by 1.2 (He/H2 = 0.2) to take the collisions with He into account (see Sect. 4.1 of the RADEX manual8 ). As the input line intensity we used the mainbeam brightness temperature, T MB . When the source is resolved, T MB is equal to the R-J equivalent radiation temperature, T R . The simulations aim to reproduce the observed line intensity and yield the values of τ0 and T ex , and the total column density of the molecule (Ntot ). We varied T kin and hn(H2 )i according to their errors to estimate the uncertainties associated with τ0 , T ex , and Ntot . The uncertainties in T MB and ∆v were not taken into account, but test calculations showed that they lead to errors in column density that are comparable to those derived from the errors in temperature and density. We also note that the use of higher H2 densities for the cores derived from the SABOCA map [Col. (4), Table 6] would lead to lower column densities of the molecules because then the excitation would be closer to thermalisation. For N2 D+ and o-D2 CO there are no molecular data files available in the LAMDA database. The line optical thicknesses and total beam-averaged column densities of these molecules were determined through LTE modelling with CLASS/Weeds. The input parameters for a Weeds model are Ntot , T ex , source size (θs ), linewidth (FWHM), and offset from the reference-channel velocity. The linewidth is directly determined from the observed line profile, so there are basically three free parameters left (Ntot , T ex , θs ). Some of the model parameters may be degenerate, and cannot be determined independently (Schilke et al. 2006; Maret et al. 2011). The source size is degenerate with excita8 Lbol b [L⊙ ] 1.2 ± 0.1 2.3 ± 0.2 0.6 ± 0.04 tion temperature in the case of completely optically thick lines (τ ≫ 1), and with column density if the lines are completely optically thin (τ ≪ 1). We assumed that the source fills the telescope beam, i.e., that the beam filling factor is unity9 , and thus the line brightness temperature is T B ≃ T MB [see Eqs. (1) and (2) in Maret et al. (2011)]. For N2 D+ , we used as T ex the values obtained for N2 H+ from RADEX simulations. For the asymmetric top rotor o-D2 CO we adopted the value T ex = 2/3×Eu /kB , which gives a lower limit to the column density (Hatchell et al. 1998). The input Ntot was then varied until a reasonable fit to the line was obtained (see Fig. 7). The associated error estimate is based on the 10% calibration uncertainty. In the derivation of column densities, the line-strength contribution of hf components within the detected lines was taken into account, i.e., the column densities were corrected for the fraction of the total line strength given in brackets in Col. (6) of Table 5. We also determined the HCO+ column density from the column density of H13 CO+ . For this calculation, it was assumed that the carbon-isotope ratio is [12 C]/[13 C] = 60 (Wilson & Rood 1994; Savage et al. 2002). This value has been used in several previous studies, e.g., by Bergin et al. (1999) in their study of dense cores in Orion (including IRAS05399)10. 9 The 27′′ beam of our N2 D+ observations is comparable to the extent of the strongest dust emission region within the cores. On the other hand, N2 D+ emission has been found to trace dust emission very well in low-mass dense cores (e.g., Crapsi et al. 2005). Therefore, the assumption of unity beam filling factor is reasonable. For o-D2 CO, the beam filling factor may be < 1. For example, Bergman et al. (2011), using the same resolution as we (27′′ ), derived the filling factor of ≃ 0.5 for the o-D2 CO(40, 4 − 30, 3 ) emission region in ρ Oph A. Moreover, the filling factor is lower if the cores contain unresolved small-scale structure. 10 Spectral lines of 12 C-isotopologue of HCO+ are likely to be optically thick. Therefore, the HCO+ deuteration can be better investigated http://www.sron.rug.nl/∼vdtak/radex/radex− manual.pdf through the DCO+ /H13 CO+ column density ratio. However, a caveat Miettinen et al.: Dense cores in Orion B9 We calculated the fractional abundances of the molecules by dividing the molecular column density by the H2 column density: x(mol) = N(mol)/N(H2 ). For this purpose, the values of N(H2 ) were derived from the LABOCA dust continuum map smoothed to the corresponding resolution of the line observations. The resolution of the H13 CO+ observations (18′′ ) is slightly better, but comparable, to that of the original LABOCA data, and thus no smoothing was done in this case. The derived column densities and abundances are listed in Table 9. We stress that the reported uncertainties are formal and optimistic, and probably underestimate the true uncertainties. A stock chart showing the fractional abundances (excluding the additional velocity components) is presented in Fig. 8. The abundance errors were derived by propagating the errors in N(mol) and N(H2 ). (undepleted) abundance, and x(CO)obs is the observed CO abundance, fD is given by fD = (2) We adopted the standard value x(CO)can = 9.5 × 10−5 (Frerking et al. 1982). To calculate the “canonical” C17 O abundance we assumed the oxygen-isotopic ratio of [16 O]/[18 O] = 500 (e.g., Wilson & Rood 1994; Williams et al. 1998). When this is combined with the [18 O]/[17 O] ratio, for which we use the standard value 3.52 (Frerking et al. 1982), the value of x(C17 O)can can be calculated as x(C17 O)can = Fig. 7. Example of the Weeds LTE modelling outlined in Sect. 4.4. The synthetic model spectrum is overlaid as a red line. x(CO)can . x(CO)obs x(CO)can x(CO)can = . 1760 [18 O]/[17 O] × [16 O]/[18 O] (3) The depletion factor fD is then calculated from fD = x(C17 O)can /x(C17 O)obs . The results are listed in Col. (2) of Table 10, where the ±-errors quoted were calculated by propagating the uncertainty in x(C17 O)obs . The degree of deuterium fractionation in HCO+ and N2 H+ was calculated by dividing the column density of the deuterated isotopologue by its normal hydrogen-bearing form as RD (HCO+ ) ≡ N(DCO+ )/N(HCO+ ) and RD (N2 H+ ) ≡ N(N2 D+ )/N(N2 H+ ). The error in RD was derived from the errors in the corresponding column densities [see Cols. (3) and (4) of Table 10]. We note that the N2 H+ and N2 D+ column densities were calculated using the non-LTE and LTE models, respectively. Although not taken into account here, this may introduce an additional error of a factor of a few in RD , which should be bear in mind. Note, however, that in all cases the normal and deuterated forms of the molecule show line emission at similar radial velocities and with similar linewidths. Therefore, the two transitions are probably tracing the same gas which makes the derived deuteration levels reasonable. 4.6. Fractional ionisation Fractional abundances of all the observed ionic species and their different isotopologues could be determined only towards IRAS05399 and SMM 4-LVC. Therefore, the fractional ionisation and cosmic-ray ionisation rate of H2 were estimated only for these two sources. A rough estimate of the lower limit to the ionisation degree can be obtained by simply summing up the abundances of the ionic species (e.g., Caselli et al. 2002a; Paper I): Fig. 8. A bar representation of fractional abundances on a logarithmic scale. The H13 CO+ and HCO+ abundances derived towards IRAS05399 are shown by a slightly stretch stock for a better illustration. 4.5. CO depletion and deuterium fractionation To estimate the amount of CO depletion in the cores, we calculated the CO depletion factor, fD . If x(CO)can is the “canonical” should be noted here. The HCO+ molecules are produced directly from CO (Sect. 4.6). On the other hand, at low temperature, CO is susceptible to the exothermic isotopic charge exchange reaction 13 C+ + 12 CO → 12 + C + 13 CO + ∆E, where ∆E/kB = 35 K (Watson et al. 1976). This is expected to cause considerable 13 C-fractionation in cold and dense gas, which complicates the deuteration analysis. x(e) > x(HCO+ ) + x(H13 CO+ ) + x(DCO+ ) + x(N2 H+ ) + x(N2 D+ ) . (4) This is based on the gas quasi-neutrality: the electron abundance equals the difference between the total abundances of the cations and anions. The resulting values are x(e) > 1.5 ×10−8 for IRAS05399, and x(e) > 6.1 × 10−8 for SMM 4-LVC. In the outer envelope where CO is not heavily depleted, HCO+ is expected to be the main molecular ion. On the other hand, we do not have observational constraints on the abundances of H+ , H+3 (and its deuterated isotopologues), H3 O+ , and metal ions (such as C+ ), all of which could play an important role in the ionisation level. When the fractional ionisation in the source is determined, the abundance ratio RH ≡ [HCO+ ]/[CO] can be used to infer the cosmic-ray ionisation rate of H2 , ζH2 . By deriving a steadystate equation for the H+3 abundance, and applying it in the corresponding equation for HCO+ , it can be shown that 13 Miettinen et al.: Dense cores in Orion B9 Table 9. Molecular column densities and fractional abundances with respect to H2 . Source IRAS 05399-0121 SMM 1 SMM 3 IRAS 05405-0117 SMM 4 2nd v-comp.b SMM 5 SMM 6 Ori B9 N 2nd v-comp.b SMM 7 IRAS 05399-0121 SMM 1 SMM 3 IRAS 05405-0117 SMM 4 2nd v-comp.b SMM 5 SMM 6 Ori B9 N 2nd v-comp.b SMM 7 N(C17 O) [1014 cm−2 ] 3.4 ± 0.5 5.6 ± 0.5 1.3 ± 0.3 1.3 ± 0.4 ... 2.0 ± 0.5 1.0 ± 0.4 1.4 ± 0.4 0.9 ± 0.3 2.3 ± 0.5 5.2 ± 0.8 x(C17 O) [10−8 ] 1.7 ± 0.4 2.9 ± 0.5 0.5 ± 0.1 1.5 ± 0.5 ... 1.5 ± 0.5 1.5 ± 0.6 1.3 ± 0.4 2.6 ± 1.0 6.7 ± 2.6 3.3 ± 1.0 N(N2 H+ ) [1013 cm−2 ] 1.5 ± 0.3 1.2 ± 0.3 0.8 ± 0.2 2.3 ± 0.3 ... 1.0 ± 0.3 0.4 ± 0.1 3.0 ± 0.3 0.1c 0.4 ± 0.1 0.5 ± 0.1 x(N2 H+ ) [10−10 ] 6.5 ± 1.9 5.4 ± 1.6 3.5 ± 1.0 23.9 ± 4.0 ... 6.2 ± 2.5 5.5 ± 1.5 22.3 ± 2.7 3.3 ± 0.6 13.4 ± 5.4 2.8 ± 0.9 N(N2 D+ )a [1012 cm−2 ] 3.1 ± 0.3 11.9 ± 1.2 2.7 ± 0.3 0.8 ± 0.1 ... 9.5 ± 1.0 ... 17.6 ± 1.8 ... ... ... x(N2 D+ ) [10−10 ] 1.5 ± 0.3 6.1 ± 1.1 1.1 ± 0.2 0.9 ± 0.2 ... 6.8 ± 2.0 ... 15.4 ± 1.9 ... ... ... N(H13 CO+ ) [1012 cm−2 ] 6.4 ± 1.0 ... ... ... ... 18.2 ± 7.8 ... ... ... ... ... x(H13 CO+ ) [10−10 ] 2.3 ± 0.6 ... ... ... ... 9.6 ± 4.8 ... ... ... ... ... N(HCO+ ) [1014 cm−2 ] 3.8 ± 0.6 ... ... ... ... 10.9 ± 4.7 ... ... ... ... ... x(HCO+ ) [10−8 ] 1.4 ± 0.4 ... ... ... ... 5.8 ± 2.9 ... ... ... ... ... N(DCO+ ) [1012 cm−2 ] 7.6 ± 0.9 13.1 ± 1.0 2.3 ± 0.3 ... ... 10.4 ± 5.0 ... 25.9 ± 2.0 1.8 ± 0.3 6.4 ± 2.0 10.0 ± 2.0 x(DCO+ ) [10−10 ] 3.6 ± 0.9 5.8 ± 0.9 1.0 ± 0.2 ... ... 6.3 ± 3.5 ... 18.7 ± 1.9 6.3 ± 1.5 22.9 ± 10.2 5.5 ± 1.8 N(o − D2 CO)a [1012 cm−2 ] 2.0 ± 0.2 1.0 ± 0.1 ... ... ... ... ... ... ... ... ... x(o − D2 CO) [10−11 ] 9.8 ± 2.2 5.1 ± 0.9 ... ... ... ... ... ... ... ... ... Notes. (a) The N2 D+ and o-D2 CO column densities were derived from Weeds models. For o-D2 CO we assumed that T ex = 2/3 × Eu /kB .(b) For SMM 4-LVC (∼ 1.6 km s−1 ) and Ori B9 N-LVC (∼ 1.9 km s−1 ), T kin = 10.4 ± 1.4 K and 13.6 ± 2.5 K, respectively (Paper II).(c) The associated error derived by varying T kin and hn(H2 )i in the RADEX calculation is negligible. RH = h [ζH2 /n(H2 )]kH+3 i, ih βH+3 x(e) + kH+3 x(CO) + kgr, H+3 x(g) βHCO+ x(e) + kgr, HCO+ x(g) (5) where kH+3 is the rate coefficient for the reaction H+3 + k H+ 3 CO −→ HCO+ + H2 , βH+3 and βHCO+ are the dissociative recombination rate coefficients of H+3 and HCO+ , and kgr, H+3 and kgr, HCO+ are the rate coefficients for the recombination of H+3 and HCO+ onto dust grains. The values of kH+3 and βHCO+ were taken from the UMIST database11 (Woodall et al. 2007), whereas βH+3 , kgr, H+3 , and kgr, HCO+ were interpolated from Pagani et al. (2009a; Tables A.1 and B.1 therein). For the grain abundance, x(g), we used the value 2.64 × 10−12 , which is based on the grain radius a = 0.1 µm, density ρgrain = 3 g cm−3 , and the dust-to-gas mass ratio Rd = 1/100 [see, e.g., Eq. (15) in Pagani et al. (2009a)]. To calculate ζH2 , we adopted as x(e) the summed abundance of ionic species. As n(H2 ) of SMM 4-LVC, we used the density of SMM 4. The obtained values are ζH2 ∼ 2.6 × 10−17 s−1 towards IRAS05399, and ∼ 4.8 × 10−16 s−1 towards SMM 4-LVC. These values should be taken as lower limits in the sense that we have used the lower limits to x(e). 5. Discussion 5.1. Dust properties By fixing the value of β to 1.9, we derived the dust temperatures in the range ∼ 7.9−10.8 K. These are about 0.5–5.5 K lower than the gas temperatures in the same objects as derived from NH3 . 11 14 http://www.udfa.net/ Table 10. The degree of CO depletion and deuterium fractionation. Source IRAS 05399-0121 SMM 1 SMM 3 IRAS 05405-0117 SMM 4 2nd v-comp. SMM 5 SMM 6 Ori B9 N 2nd v-comp. SMM 7 fD 3.2 ± 0.7 1.9 ± 0.3 10.8 ± 2.2 3.6 ± 1.2 ... 3.6 ± 1.2 3.6 ± 1.4 4.2 ± 1.3 2.1 ± 0.8 0.8 ± 0.3 1.6 ± 0.5 RD (HCO+ ) 0.020 ± 0.004 ... ... ... ... 0.010 ± 0.006 ... ... ... ... ... RD (N2 H+ ) 0.207 ± 0.046 0.992 ± 0.267 0.338 ± 0.092 0.035 ± 0.006 ... 0.950 ± 0.302 ... 0.587 ± 0.084 ... ... ... Due to the large uncertainties associated with T dust , the near equality T kin ≃ T dust seems possible, as expected at high densities where collisional coupling between the gas and dust becomes efficient. Theoretical models have shown that in the dense interiors of starless cores [n(H2 ) & 3 × 104 cm−3 ], the gas and dust temperatures are similar, although the gas can be slightly warmer due to cosmic-ray heating (e.g., Galli et al. 2002). Some of the dust temperatures we derived are very low. In particular, for SMM 6 we obtained T dust ≃ 7.9+3.0 −1.8 K. Theoretical models (e.g., Evans et al. 2001) and previous observational studies (e.g., Schnee et al. 2007a; Crapsi et al. 2007; Harju et al. 2008) have indicated that very low gas and dust temperatures (∼ 6 − 7 K) can be reached in the dense interiors of dense cores. The dust emissivity spectral indices we derived, β ∼ 0.5 − 1.8, are physically reasonable, and suggest that the as- Miettinen et al.: Dense cores in Orion B9 sumption T dust = T kin is valid (cf. Schnee & Goodman 2005). Furthermore, the β values derived towards SMM 3, 5, and 7 are close to 1.9. We note that the dense gas and dust associated with the additional velocity components along the line of sight also affect the observed dust continuum properties (e.g., SMM 4). Therefore, the derived β values for these sources should be taken with caution. We note, however, that a decrease of β from the ’fiducial’ value 2 to a shallower emissivity spectral index could be related to dust grain coagulation in the inner parts of dense cores (e.g., Miyake & Nakagawa 1993; OH94). Observational studies have found evidence that β decreases as a result of grain growth at high densities (e.g., Goldsmith et al. 1997; Visser et al. 1998). More recently, Kwon et al. (2009) found that β . 1 for their sample of three Class 0 sources, resembling the values we found towards the Class 0 candidates IRAS05405 and SMM 4. These results suggest that dust grains in the envelopes of Class 0 protostars can grow in size leading to a shallower spectral index of dust emissivity. Based on Herschel observations of cold interstellar clouds detected with the Planck satellite, Juvela et al. (2011) found that β decreases down to ∼ 1 near internal heating sources. Radiative transfer modelling, however, suggests that such a decrement is due to line-of-sight temperature variations rather than changes in the grain properties. On the other hand, some studies of low-mass dense cores suggest that β could be larger than 2 (e.g., Shirley et al. 2005; Schnee et al. 2010a; Shirley et al. 2011). Also, recent studies of Planck-detected cold Galactic clumps indicate that β > 2 (Planck Collaboration et al. 2011). It is possible that β is anticorrelated with T dust , so that in cold, dense cores the emissivity spectral index is steeper than 2. 5.2. Refined SEDs Inclusion of the new 350 µm data to the source SEDs confirmed our earlier protostellar classifications. In Paper I, we classified the sources SMM 3, SMM 4, and IRAS05405 as Class 0 objects. For such objects, the bolometric temperature is T bol < 70 K, Lsubmm /Lbol ratio is > 0.005, and they are charac0.6 terised by the ratio Menv /L0.6 bol & 0.4 M⊙ /L⊙ (André et al. 1993; Bontemps et al. 1996; André et al. 2000). All these conditions are fulfilled after the refined SED analysis. The SED models presented in Paper I suggested 350-µm flux densities of 21.7 Jy for SMM 3, 13.5 Jy for IRAS05405, and 11.1 Jy for SMM 4 – much higher than determined in the present work. Compared to the SED results in Paper I, the envelope mass of SMM 3 and 4 is 1.0 M⊙ higher and 1.8 M⊙ lower, respectively; the mass of IRAS05405 remains the same. The refined luminosity is lower by a factor of ∼ 3 in all three cases. The temperature of the envelope, T cold , resulting from the new SEDs is only 8 K for all sources. This is much lower than the values 11.6–16.1 K derived in Paper I. The Lcold /Lbol ratios derived here are also much smaller than estimated previously. Unlike deduced in Paper I, the cold component does not appear to dominate the total source luminosity. The Lsubmm /Lbol and Menv /L0.6 bol ratios derived here are mostly comparable to those obtained in Paper I. The angular resolution of the Spitzer images used is about 6′′ at 24 µm, and 18′′ at 70 µm. Photometry was done by pointsource fitting, and the aperture size for photometry was 13′′ at 24 µm, and 35′′ at 70 µm (Paper I). The submm data points used in the SEDs were obtained at about 20′′ resolution within a 40′′ aperture. Therefore, the 70, 350, and 870-µm flux densities refer to a similar spatial scale (∼ 0.09 pc). Direct comparison of these data with the 24 µm data is reasonable because we are dealing with 24-µm point sources. Our model SEDs therefore produce an approximation to the core parameters on ∼ 0.09 pc spatial scale. Note that the sources SMM 4 and 4b are treated as one source and thus the corresponding SED should be taken with caution (Sect. 5.7.2). 5.3. Molecular column densities and abundances The present N2 H+ column densities are about 0.6–5 times the corresponding LTE column densities presented in Paper II; within the errors, the two values are comparable in the case of IRAS05399, SMM 1, 3, and 7. Bergin et al. (1999) found the column densities N(N2 H+ ) ∼ 4.0 ± 0.3 × 1012 cm−2 , N(H13 CO+ ) ∼ 6.9 ± 1.3 × 1011 cm−2 , and N(DCO+ ) ∼ 1.4 ± 0.2 × 1012 cm−2 towards IRAS05399 using a statistical equilibrium model. Our values, which are derived towards a position of about 12′′ northeast from the target position of Bergin et al. (1999), are about 3.8 ± 0.8, 9.3 ± 2.3, and 5.4 ± 1.0 times higher, respectively. Moreover, the C18 O column density of ∼ 5.8 ± 0.2 × 1015 cm−2 derived by Bergin et al. (1999) towards IRAS05399 implies the value N(C17 O) ∼ 1.6 ± 0.1 × 1015 cm−2 . The latter value is 4.7 ± 0.3 times higher than the value we have obtained. Besides a slightly different target position, these discrepancies are likely caused by the fact that Bergin et al. (1999) assumed optically thin emission, and the values n(H2 ) ∼ 105 cm−3 and T kin = 15 K in their non-LTE excitation analysis. The C17 O column densities and abundances we have derived are comparable to those found by Bacmann et al. (2002) for their sample of seven prestellar cores, and by Schnee et al. (2007b) towards the starless core TMC1C. The N2 H+ column densities are also comparable to those determined towards several other low-mass starless cores and Class 0 protostellar objects (Caselli et al. 2002b; Crapsi et al. 2005; Roberts & Millar 2007; Daniel et al. 2007; Emprechtinger et al. 2009; Friesen et al. 2010a,b). On the other hand, the N2 D+ column densities we have obtained are generally larger than those derived for other low-mass cores in the studies mentioned above. Also, the H13 CO+ and DCO+ column densities and abundances we derive are higher than those obtained by Anderson et al. (1999) in the R CrA region, and by Frau et al. (2010) for a sample of starless cores in the Pipe Nebula. The HCO+ abundance of ∼ 1.4 × 10−8 we have derived towards IRAS05399 is relatively high. HCO+ could, in principle, increase in abundance at later evolutionary stages, when the CO gas-phase abundance is enhanced, and the reaction H+3 + CO → HCO+ + H2 becomes efficient. However, the derived CO depletion factor in IRAS05399 is relatively large, fD ∼ 3.2, and thus the high HCO+ abundance may be the result of an alternative production pathway, C+ + H2 O → HCO+ + H, viable in the shocks (Rawlings et al. 2000; Viti et al. 2002). We note that IRAS05399 drives the HH92 jet, which lies close to the plane of the sky, and the presence of shock is therefore plausible (Gredel et al. 1992; Bally et al. 2002). 5.4. Depletion and deuteration The CO depletion factors we have derived are in the range fD ∼ 1.6 ± 0.5 − 10.8 ± 2.2 ( fD = 3.6 ± 1.2 for SMM4-LVC and 0.8 ± 0.3 for Ori B9 N-LVC). Interestingly, the strongest depletion is observed towards the protostellar core SMM 3, and the lowest fD value is found in the starless core SMM 7. This could be caused by the fact that we do not observe exactly towards the submm 15 Miettinen et al.: Dense cores in Orion B9 peak of SMM 7, whereas we probe the dense envelope of SMM 3 (see Figs. 1 and 2). The second lowest fD value, 1.9 ± 0.3, is seen towards the starless core SMM 1. Bergin et al. (1999) mapped the IRAS05399/SMM 1-system in C18 O, CS, H13 CO+ , and DCO+ , and found that the emission peaks coincide with the position of SMM 1, in agreement with our finding of low CO depletion. In general, the gas-phase CO abundance is expected to decrease during the prestellar phase of core evolution as a result of freeze out onto grain surfaces. During the protostellar phase, on the other hand, CO is expected to be released back into the gas phase via ice-mantle sublimation in the warmer envelope surrounding the protostar. Caselli et al. (2008) estimated the value fD (CO) = 3.6 towards a position at the edge of Ori B9 N using the data from Caselli & Myers (1995) and Harju et al. (2006). This is higher than the value 2.1 ± 0.8 we have derived towards our target position near Ori B9 N (also at the core edge). For comparison, Bacmann et al. (2002) studied the level of CO depletion in prestellar cores, and found the values in the range 4.5–15.5. These are comparable to the values we have obtained. Moreover, Bacmann et al. (2002) found a positive correlation between fD and n(H2 ) of the form fD ∝ n(H2 )0.4−0.8 . No correlation was found between fD and n(H2 ) in the present study. Emprechtinger et al. (2009) derived the values fD ∼ 0.3 − 4.4 towards a sample of Class 0 protostars. We note that in the depletion analysis we used the value 9.5× 10−5 for the undepleted CO abundance. However, this value is known to vary by a factor of ∼ 2 − 3 between different starforming regions. For example, Lacy et al. (1994) determined the CO abundance of ∼ 2.7 × 10−4 towards NGC 2024 in Orion B. Adopting the latter value would result in ∼ 2.8 times larger depletion factors. The N2 H+ deuteration degree in the Orion B9 cores is found to be in the range RD (N2 H+ ) ≃ 0.035 − 0.992. The extreme value of 0.992 measured towards SMM 1 is, to our knowledge, the highest level of deuteration in N2 H+ observed so far. This suggests that the core is chemically highly evolved but is in contradiction with the low fD value which points towards a younger evolutionary stage. SMM 1 could have been affected by the outflow driven by IRAS05399, releasing CO into the gas phase and effectively resetting the chemical clock. In this case, the very high RD (N2 H+ ) value would be remnant of the earlier CO-depleted phase, and not yet affected by the gas-phase CO which destroys N2 H+ and N2 D+ . An opposite situation was found by Crapsi et al. (2004) towards the chemically evolved dense core L1521F which harbours a very low luminosity object or VeLLO (Bourke et al. 2006), where fD ∼ 15 but RD (N2 H+ ) is only ∼ 0.1. For comparison, Crapsi et al. (2005) derived the values of RD (N2 H+ ) in the range . 0.02 − 0.44 for their sample of starless cores, and Daniel et al. (2007) derived comparable values of RD (N2 H+ ) ∼ 0.07 − 0.53 towards another sample of starless cores. Also, Roberts & Millar (2007) found RD (N2 H+ ) values of < 0.01 − 0.31 for low-mass cores. Emprechtinger et al. (2009) found the values RD (N2 H+ ) < 0.029 − 0.271 towards Class 0 objects. Pagani et al. (2009a) and Fontani et al. (2011) derived the value RD (N2 H+ ) ∼ 0.7 at the centre of the starless core L183 and towards the high-mass prestellar core candidate G034-G2 MM2, respectively, which were the highest fractionations reported before the present work. The degree of deuterium fractionation in HCO+ could be derived only towards IRAS05399 and SMM 4-LVC with the values RD (HCO+ ) ≃ 0.020 and 0.010, respectively. These are significantly lower than the corresponding RD (N2 H+ ) values, in agreement with the results by Roberts & Millar (2007) 16 and Emprechtinger et al. (2009). Such a trend is believed to be caused by the role of CO, which is the parent species of HCO+ and DCO+ , but the main destroyer of H2 D+ , N2 H+ , and N2 D+ molecules. Therefore, molecular deuteration proceeds most efficiently in regions where CO is depleted. In the warmer envelope layers where CO is not depleted, HCO+ can have a relatively high abundance, resulting in a lower line-ofsight average value of RD (HCO+ ) compared to that of N2 H+ (Emprechtinger et al. 2009). Note that we have employed the DCO+ (4 − 3) transition, which is expected to trace the inner part of the high-density envelope, where heating by the embedded protostar may affect the deuterium chemistry [by lowering the RD (HCO+ ) ratio]. For comparison, Jørgensen et al. (2004) determined similar RD (HCO+ ) values of . 0.001 − 0.05 for a sample of 18 low-mass pre- and protostellar cores. Previous studies of low-mass starless and protostellar cores have found correlations between the degree of CO depletion and deuterium fractionation (Bacmann et al. 2003; Jørgensen et al. 2004; Crapsi et al. 2005; Emprechtinger et al. 2009). No correlation was found between fD and RD (N2 H+ ) in the present study. However, the values of fD and RD (N2 H+ ) for the protostellar cores SMM 3 and IRAS05399 are in rough agreement with the finding of Emprechtinger et al. (2009), i.e., that deuteration in the envelopes of Class 0 protostars increases as a function of fD . 5.5. Fractional ionisation In Paper I, we determined the lower limits to x(e) of a few times 10−8 , and upper limits of about six times 10−7 towards two target positions near the clump associated with IRAS05405. The lower limits to x(e) derived in the present work are comparable to those from Paper I. The standard relation between the electron abundance and the H2 number density is x(e) ∼ 1.3 × 10−5 n(H2 )−1/2 (McKee 1989). This is based on the pure cosmic-ray ionisation with the rate 1.3 × 10−17 s−1 and includes no depletion of heavy elements. The standard relation yields the values x(e) ≃ 5.5 × 10−8 for IRAS05399, and ≃ 6.7 × 10−8 for SMM 4-LVC. These are roughly comparable to the estimated lower limits to x(e). The values of ζH2 were found in Paper I to be ∼ 1−2×10−16 s−1 . These resemble the value derived here towards SMM 4LVC, but are an order of magnitude higher than obtained for IRAS05399. Instead, the value ζH2 ∼ 2.6 × 10−17 s−1 derived towards IRAS05399 is within a factor of two of the standard value. For comparision, observations of the Horsehead Nebula in Orion B by Goicoechea et al. (2009) could only be reproduced with ζH2 = 7.7 ± 4.6 × 10−17 s−1 . Caselli et al. (1998) determined x(e) in a sample of 24 lowmass cores consisting of both starless and protostellar objects. Their analysis was based on observations of CO, HCO+ , and DCO+ , and the resulting values were in the range 10−8 to 10−6 , bracketing the values for the Orion B9 cores. They argued that the variation in x(e) among the sources is due to variations in metal abundance and ζH2 ; the latter was found to span a range of two orders of magnitude between 10−18 − 10−16 s−1 . Some of this variation could be due to different cosmic-ray flux in the source regions. Williams et al. (1998) used observations of C18 O, H13 CO+ , and DCO+ to determine the values 10−7.5 . x(e) . 10−6.5 in a similar sample of low-mass cores as Caselli et al. (1998), but using a slightly different analysis. They deduced a mean value of ζH2 = 5×10−17 s−1 . Applying the same analysis as Williams et al. (1998), Bergin et al. (1999) found the ionisation levels of 10−7.3 . x(e) . 10−6.9 towards more massive cores Miettinen et al.: Dense cores in Orion B9 (in Orion) than to those studied by Williams et al.. Bergin et al. (1999) determined the fractional ionisation towards IRAS05399 to lie in the range x(e) ∼ 9.3 × 10−8 − 1.8 × 10−7. The lower limit we have derived, x(e) ∼ 1.5 × 10−8, is about six times lower than the corresponding value derived by Bergin et al. (1999). Recently, Padovani & Galli (2011) suggested that the magnetic field threading the core affects the penetration of cosmic rays, and can decrease the ionisation rate by a factor of ∼ 3 − 4. The values of ζH2 determined through observations (so far) would then underestimate the inter-core values by the above factor, making the ζH2 values more comparable with those measured in diffuse clouds. 5.6. Deuterated formaldehyde We have detected the ortho form of doubly deuterated formaldehyde, D2 CO, towards the prestellar core SMM 1 and the protostellar core IRAS05399. We derived the column densities and abundances of this molecule to be 1 − 2 × 1012 cm−2 and ∼ 5 − 10 × 10−11 , respectively. We note that the high-temperature statistical ortho/para ratio of D2 CO is 2, and appears to be similar even in cold star-forming cores (see Roberts & Millar 2007). Using the IRAM 30-m telescope, Ceccarelli et al. (1998) detected D2 CO towards IRAS 16293-2422. It was the first reported detection of D2 CO towards a low-mass star-forming core. A lower limit to the D2 CO column density they obtained, ∼ 1014 cm−2 , is much higher than in our sources. Ceccarelli et al. (1998) speculated that such a high D2 CO column density requires evaporation of D2 CO from the grain mantles, where it is expected to be formed during the prestellar phase. Later, Ceccarelli et al. (2001) demonstrated that heating by the central protostar in IRAS 16293-2422 is responsible for the mantle evaporation, and injection of D2 CO into the gas phase. Loinard et al. (2002) detected D2 CO towards 19 low-mass protostellar cores, and found the D2 CO/H2 CO abundance ratios of ∼ 0.02 − 0.4. Bacmann et al. (2003) and Roberts & Millar (2007) found D2 CO column densities of ∼ 0.5 − 2.7 × 1012 cm−2 towards a sample of lowmass prestellar and protostellar cores. These column densities are similar to those we have derived. However, the D2 CO abundances derived by Bacmann et al. (2003) are somewhat lower than those we have derived. Recently, Bergman et al. (2011) found the D2 CO column densities of ∼ 2 × 1012 cm−2 towards a few positions in ρ Oph A, which are very similar to our values; towards the D-peak of ρ Oph A, however, they derived a high value of ∼ 3.2 × 1013 cm−2 . D2 CO is expected to evaporate from the CO-rich grain mantles when the dust temperature exceeds about 25 K (Ceccarelli et al. 2001). This would agree with the low CO depletion factor of 1.9 derived towards SMM 1, because the CO sublimation temperature is about 20 K (e.g., Aikawa et al. 2008). It is uncertain, however, why we see higher degree of CO depletion towards the nearby protostellar core IRAS05399. Moreover, the gas kinetic temperature is only 11.9 K in SMM 1 and 13.5 K in IRAS05399. It is possible, that the presence of gas-phase D2 CO in these two sources is due to a non-thermal desorption mechanism (Roberts & Millar 2007). Bergman et al. (2011) suggested that cosmic-ray heating and the formation energy of the newly formed species could play an important role in the release of D2 CO from the grain mantles in starless cores. In addition to these two mechanisms, shocks caused by the protostellar outflow driven by IRAS05399 could be responsible in releasing D2 CO from the icy grain mantles in the IRAS05399/SMM 1-region. Towards IRAS05399 the emission is perhaps dominated by the cool envelope whereas in SMM 1 we might see a component which is interacting with the outflow (causing the high gas-phase CO abundance). 5.7. Fragmentation in the Orion B9 cores 5.7.1. SMM 6 – a fragmented prestellar core Figure 9 shows the SABOCA 350-µm image of the prestellar core SMM 6 overlaid with the LABOCA 870-µm contours. The core is filamentary in shape, and it is resolved into three to four subcondensations at 350 µm. The projected linear extent of core’s long axis is about 1.′ 9 or 0.25 pc, and the core’s massper-length is Mline ≃ 33 M⊙ pc−1 . The length-to-width ratio increases from about 1.9 at the NW end to about 7.4 at the SE end. The projected separation between the condensations is ∼ 29′′ or ∼ 0.06 pc, where the sources 6c and 6d are treated as a single fragment. The measured separations should be taken as lower limits because of the possible projection effects. To analyse the fragmentation of SMM 6 in more detail, we 2 calculate p its thermal Jeans length from λJ = cs /(GΣ0 ), where cs = kB T kin /µmH is the isothermal sound speed (at T kin = 11 K), G is the gravitational constant, and Σ0 = µmH N(H2 ) is the surface density; µ is the mean molecular weight per free particle (2.33 for He/H= 0.1), and N(H2 ) refers to the central column density for which we use the value ∼ 1022 cm−2 [see Col. (3) of Table 6]. The resulting Jeans length is ∼ 0.05 pc, very similar to the observed separation between the condensations. In Paper II, we found from NH3 measurements that turbulent motions within SMM 6 are subsonic, so their contribution to the effective sound speed would not increase the calculated Jeans length much. Assuming spherical geometry, the local Jeans mass of SMM 6 is MJ = 4π/3 × hρi(λJ /2)3 ∼ 2.2 M⊙ . The corresponding Jeans number is nJ = M/MJ ∼ 4, which is similar to the number of observed subfragments. If the core substructure is caused by gravitational fragmentation, nJ is expected to give an approximate number of subfragments within the core (e.g., Rathborne et al. 2007). The measured radii of the individual condensations, ∼ 0.01− 0.02 pc, and their masses, p ∼ 0.1 − 0.2 M⊙ , are smaller than the Jeans lengths of λJ = πc2s /Gρ ∼ 0.05 − 0.06 pc12 , and Jeans masses of ∼ 0.8 − 0.9 M⊙ [assuming T kin = 11 K and densities from Col. (4) of Table 6]. Because the subcondensations are presumably colder than 11 K, their masses and densities are likely to be higher. For instance, using the dust temperature T dust = 7.9 K derived earlier for SMM 6 would result in about 4.4 times higher masses and densities. Lower temperature would also cause the Jeans length and mass to be smaller. At 7.9 K, these would be ∼ 0.02 pc and ∼ 0.2 − 0.3 M⊙ , respectively. Therefore, the observed condensation properties can well be comparable to the corresponding Jeans values. The above analysis suggests that thermal Jeans instability is the dominant process responsible for the core fragmentation. Moreover, the process of Jeans-fragmentation appears to have reached its final state at scale of the observed subcondensations. Thus, the condensations are potential sites to give birth to individual stars or small stellar systems. The condensations can grow in mass through (competitive) accretion from the parent core. In general, core fragmentation is believed to be the principal mechanism for the formation of binary and multiple stellar systems (e.g., Tohline 2002; Goodwin et al. 2007). 12 For the condensations we utilise the λJ -formula which assumes spherical symmetry. For the filamentary parent core, we calculated the central Jeans length λJ ∝ 1/Σ0 . 17 Miettinen et al.: Dense cores in Orion B9 Besides the Jeans analysis, it is interesting to examine whether SMM 6, owing to its filamentary shape, is unstable to axisymmetric perturbations. For an unmagnetised isothermal filament the instability is reached if its Mline value exceeds the crit critical equilibrium value of Mline = 2c2s /G (e.g., Ostriker 1964; crit Inutsuka & Miyama 1992). For SMM 6, Mline ≈ 18 M⊙ pc−1 , crit and Mline /Mline ≈ 1.8. Thus, SMM 6 appears to be a thermally supercritical filament susceptible to fragmentation, in agreement with the detected substructure. We note that in the crit case of a magnetised molecular-cloud filament, Mline differs by only a factor of order unity from that of an unmagnetised case (Fiege & Pudritz 2000). The three-dimensional velocity dispersion in SMM 6 is σ3D = 0.384 km s−1 (from the NH3 data in Paper II). This can be used to estimate the perturbation timescale, or the signal crossing time, τcross ≡ D/σ3D , where the diameter is D = 0.25 pc for SMM 6. The value of τcross is about 6.3 × 105 yr or ∼ 2.8 times the free-fall timescale, τff . This timescale is comparable to the lifetime of the prestellar phase of core evolution of a few times τff (see Papers I and II and references therein). The observed substructure within SMM 6 shows that core fragmentation can take place already during the early prestellar phase of evolution, i.e., before the formation of embedded protostar(s). Recently, Chen & Arce (2010) suggested that the three subcondensations in the prestellar core R CrA SMM 1A were formed through the fragmentation of the elongated parent core in the isothermal phase. It seems likely that SMM 6 has undergone a similar fragmentation process. The universality of fragmentation of the starless cores is unclear, however. For example, Schnee et al. (2010b) found that none of their 11 starless cores in Perseus are fragmented into smaller subunits. This is important knowledge when comparing the core mass function (CMF) to the stellar initial mass function (IMF), because the presence of substructure is believed to be one reason for the mass shift between the CMF and the IMF. from the protostar position. Next to that, at the borderline of the 3.3σ LABOCA 870-µm contour, there is another subcondensation, SMM 3c. The projected separation between SMM 3 and 3b is about 0.08 pc. This is very close to the thermal Jeans length for the whole SMM 3 core of 0.07 pc. In SMM 7, the subcondensation SMM 7b is at ∼ 26′′ or 0.06 pc from the ’main’ submm peak. Again, this is quite close to the local Jeans length of 0.09 pc. It thus seems possible, that the substructure within SMM 3 and 7 is caused by thermal Jeans fragmentation. We note that the SMM 3/3b-system is qualitatively similar to L1448 IRS2/2E, where there is a dust condensation (IRS 2E) next to a Class 0 protostar; L1448 IRS2E is the most promising candidate of the “first hydrostatic core” detected so far (Chen et al. 2010). SMM 4 is resolved into two fragments at 350 µm. The eastern one, SMM 4b, is associated with a Spitzer source at 24 and 70 µm. In Papers I and II, we proposed that N2 H+ could be depleted in the dense envelope of SMM 4. This was based on two observational results: i) the N2 H+ (1 − 0) map of Caselli & Myers (1994) shows no emission peaks near SMM 4; and ii) the N2 H+ (3 − 2) line was not detected towards SMM 4 at the systemic velocity ∼ 9 km s−1 (see also Fig. 3). The NH3 lines, on the other hand, can be seen at ∼ 9 km s−1 with an additional velocity component at about 1.6 km s−1 (Paper II). The present molecular-line observations, however, revealed a somewhat surprising result: all the observed lines are at an LSR velocity of about 1.5–1.7 km s−1 , not at ∼ 9 km s−1 . The absence of molecular-line emission at ∼ 9 km s−1 is probably not due to chemistry, because it is not reasonable to think that all the observed species, in particular DCO+ and N2 D+ , would have been depleted (e.g., Lee et al. 2003). Instead, it seems that SMM 4 is a member of the low-velocity part of Orion B discussed in Paper II. Such a chance line-of-sight alignment is surprising, because the nearby sources SMM 5, IRAS05405, and Ori B9 N show line emission at ∼ 9 km s−1 . The reason why we detected the 9km s−1 NH3 lines towards SMM 4 could be due to the large beam size (40′′ ) of the observations; NH3 emission could have been captured from the nearby cores by the beam. Also, the morphology of the N2 H+ integrated intensity map of Caselli & Myers (1994) can be understood if the velocity range used to construct the map only covers emission near 9 km s−1 ; SMM 4 emits at lower velocity and does not show up on the map. The dust properties of SMM 4, such as its SED, should be taken with caution as the core consists of subcondensations. It also seems possible that the emission from the clump near SMM 4 is ’contaminated’ by an unrelated object(s) along the line of sight. 6. Summary and conclusions Fig. 9. Detailed 350-µm SABOCA image of the fragmented prestellar core SMM 6 with power-law scaling. The contours are as in Fig. 2. The green plus sign indicates the position of our molecular-line observations. The colour-bar scale corresponds to Jy beam−1 . 5.7.2. SMM 3, 4, and 7 The SABOCA 350-µm images of SMM 3, 4, and 7 overlaid with 870-µm contours are shown in more detail in Fig. 10. In SMM 3, there is a subcondensation, we called SMM 3b, at about 36′′ 18 We have carried out a (sub)mm study of dense cores in Orion B9. We used APEX to map the region at 350 µm, and to observe the transitions of C17 O, H13 CO+ , DCO+ , N2 H+ , and N2 D+ . These data were compared with our previous LABOCA 870-µm data. The principal aim of this study was to investigate the dust emission of the cores near the peak of their SEDs, and the degrees of CO depletion, deuterium fractionation, and ionisation in the sources. Our primary results are summarised as follows: 1. All the 870-µm cores within the boundaries of our SABOCA map were detected at 350 µm. The strongest 350-µm source in the region is SMM 3, a candidate Class-0 protostellar core. 2. Four of the 870-µm cores, namely SMM 3, 4, 6, and 7, were resolved into at least two condensations at 350 µm. In particular, the elongated prestellar core SMM 6 was resolved into three to four very low-mass subcondensations, showing that Miettinen et al.: Dense cores in Orion B9 4. We refined some of the protostellar SEDs presented in Paper I by adding the observed 350-µm flux densities. No radical changes were found, and the sources IRAS 05405-0117 and SMM 3 and 4 are classified as Class 0 objects, in agreement with our previous results. 5. The CO depletion factors were found to lie in the range fD ∼ 1.6±0.5−10.8±2.2. We found no systematic difference in fD between the starless and protostellar cores. In accordance with previous observations and theoretical predictions the most severe CO depletion is seen towards the core with highest average density, SMM 3. The degree of deuteration in N2 H+ was found to be in the range N(N2 D+ )/N(N2 H+ ) ≃ 0.035 ± 0.006 − 0.992 ± 0.267. The N(DCO+ )/N(HCO+ ) ratio was found to be about 1–2%, comparable to those seen in other low-mass star-forming regions. 6. The fractional ionisation could only be derived towards IRAS 05399-0121 and SMM 4-LVC with the lower limits of x(e) > 1.5 × 10−8 and > 6.1 × 10−8 , respectively. These values are comparable to the fractional ionisations we derived earlier towards a few target positions near IRAS 05405-0117 and SMM 4. The cosmic-ray ionisation rate of H2 implied by the derived lower x(e) limit is ζH2 ∼ 2.6 × 10−17 s−1 towards IRAS 05399-0121, and ∼ 4.8 × 10−16 s−1 towards SMM 4-LVC. The former value, which does not suffer from line-of-sight contamination, is within a factor of two of the ’standard’ value 1.3 × 10−17 s−1 . 7. The highest degree of deuteration in N2 H+ , ∼ 0.99, was derived towards the prestellar core SMM 1. To our knowledge, this is the most extreme level of N2 H+ deuteration reported so far. We also detected D2 CO emission towards SMM 1 with an abundance of several times 10−11 . Because D2 CO is expected to be formed through the grain-surface chemistry, its presence in the gas phase in SMM 1 could be due to shocks driven by the jet from IRAS 05399-0121, resulting in the release of D2 CO from the grain mantles. This conforms to the low CO depletion factor of 1.9 derived towards SMM 1. The very high N2 H+ deuteration could be remnant of the earlier CO-depleted phase, and not yet affected by the destroying effect of gas-phase CO. 8. It seems likely that the elongated clump associated with IRAS 05405-0117 and SMM 4 consists of physically independent objects that form a single clump only in projection along the line of sight. Fig. 10. Zoom-in views of Fig. 2 with linear scaling towards SMM 3, 4, and 7. The box symbols mark the positions of the Spitzer 24-µm point sources. The green plus signs show the positions of our molecular-line observations. In each panel, the colour-bar scale corresponds to Jy beam−1 . core fragmentation can take place during the prestellar phase of evolution. In all cases, the origin of substructure can be explained by thermal Jeans-type fragmentation. 3. The dust temperatures derived from the 350-to-870-µm flux +5.7 density ratio are very low, only T dust ≈ 7.9+3.0 −1.8 − 10.8−2.6 K. The corresponding gas temperatures are typically a few kelvins higher. We also derived the submm dust emissivity spectral indices using the assumption T dust = T gas, and they are in the range β ≈ 0.5 ± 0.8 − 1.8 ± 0.6. The uncertainties in β are large, and within the errors the values are comparable to the fiducial value β = 2. In Papers I and II we discussed the origin of dense cores in Orion B9. The region could have been affected by the feedback from the nearby Ori OB1b subgroup of the Orion OB1 association [resembling the process termed “cloud shuffling” by Elmegreen (1979)]. This feedback process may have played a role in sweeping up the gas into dense shells and filaments out of which the dense cores were later fragmented. This picture is supported by the kinematics of Orion B9 (low-velocity lineof-sight components). The present observational results suggest that the further fragmentation of cores into smaller condensations is caused by the thermal Jeans instability. The observations presented in this paper also reveal the intricate chemistry of the cores, such as heterogeneous depletion and deuteration among the cores. This suggests that the individual Orion-B9 cores, though possibly formed collectively in the parent filaments, are evolving chemically at different rates. Acknowledgements. We acknowledge the staff at the APEX telescope for performing the service-mode observations presented in this paper. We would also like to thank the people who maintain the CDMS and JPL molecular spectroscopy databases, and the Splatalogue Database for Astronomical 19 Miettinen et al.: Dense cores in Orion B9 Spectroscopy. The authors would like to thank the anonymous referee for useful comments. We acknowledge support from the Academy of Finland through grants 127015 and 132291. This work makes use of observations made with the Spitzer Space Telescope, which is operated by the Jet Propulsion Laboratory (JPL), California Institute of Technology under a contract with NASA. Moreover, this research has made use of NASA’s Astrophysics Data System and the NASA/IPAC Infrared Science Archive, which is operated by the JPL, California Institute of Technology, under contract with the NASA. References Aikawa, Y., Wakelam, V., Garrod, R. T., and Herbst, E. 2008, ApJ, 674, 984 Anderson, I. M., Caselli, P., Haikala, L. K., and Harju, J. 1999, A&A, 347, 983 André, Ph., Ward-Thompson, D., and Barsony, M. 1993, ApJ, 406, 122 André, Ph., Ward-Thompson, D., and Barsony, M. 2000, in Protostars and Planets IV, ed. V. Mannings, A. P. Boss, & S. S. Russell (Tuscon, AZ: Univ. Arizona Press), p. 59 Bacmann, A., Lefloch, B., Ceccarelli, C., et al. 2002, A&A, 389, L6 Bacmann, A., Lefloch, B., Ceccarelli, C., et al. 2003, ApJ, 585, L55 Bally, J., Reipurth, B., and Aspin, C. 2002, ApJ, 574, L79 Belitsky, V., Lapkin, I., Vassilev, V., et al. 2007, in Proceedings of joint 32nd International Conference on Infrared Millimeter Waves and 15th International Conference on Terahertz Electronics, September 3-7, 2007, City Hall, Cardiff, Wales, UK, pp. 326-328 Bergin, E. A., Plume, R., Williams, J. P., and Myers, P. C. 1999, ApJ, 512, 724 Bergman, P., Parise, B., Liseau, R., and Larsson, B. 2011, A&A, 527, A39 Bontemps, S., André, Ph., Terebey, S., and Cabrit, S. 1996, A&A, 311, 858 Bourke, T. L., Myers, P. C., Evans, N. J., II, et al. 2006, ApJ, 649, L37 Caselli, P., and Myers, P. C. 1994, in Clouds, Cores, and Low Mass Stars, ed. D. P. Clemens, & R. Barvainis, 65, p. 52 Caselli, P., and Myers, P. C. 1995, ApJ, 446, 665 Caselli, P., and Dore, L. 2005, A&A, 433, 1145 Caselli, P., Walmsley, C. M., Terzieva, R., and Herbst, E. 1998, ApJ, 499, 234 Caselli, P., Walmsley, C. M., Zucconi, A., et al. 2002a, ApJ, 565, 344 Caselli, P., Benson, P. J., Myers, P. C., and Tafalla, M. 2002b, ApJ, 572, 238 Caselli, P., Vastel, C., Ceccarelli, C., et al. 2008, A&A, 492, 703 Ceccarelli, C., Castets, A., Loinard, L., et al. 1998, A&A, 338, L43 Ceccarelli, C., Loinard, L., Castets, A., et al. 2001, A&A, 372, 998 Chen, X., and Arce, H. G. 2010, ApJ, 720, L169 Chen, X., Arce, H. G., Zhang, Q., et al. 2010, ApJ, 715, 1344 Crapsi, A., Caselli, P., Walmsley, C. M., et al. 2004, A&A, 420, 957 Crapsi, A., Caselli, P., Walmsley, C. M., et al. 2005, ApJ, 619, 379 Crapsi, A., Caselli, P., Walmsley, M. C., and Tafalla, M. 2007, A&A, 470, 221 Daniel, F., Cernicharo, J., Roueff, E., et al. 2007, ApJ, 667, 980 Elmegreen, B. G. 1979, ApJ, 231, 372 Emprechtinger, M., Caselli, P., Volgenau, N. H., et al. 2009, A&A, 493, 89 Evans, N. J., II, Rawlings, J. M. C., Shirley, Y. L., and Mundy, L. G. 2001, ApJ, 557, 193 Fiege, J. D., and Pudritz, R. E. 2000, MNRAS, 311, 85 Fontani, F., Palau, A., Caselli, P., et al. 2011, A&A, 529, L7 Frau, P., Girart, J. M., Beltrán, M. T., et al. 2010, ApJ, 723, 1665 Frerking, M. A., Langer, W. D., and Wilson, R. W. 1982, ApJ, 262, 590 Friesen, R. K., Di Francesco, J., Shimajiri, Y., & Takakuwa, S. 2010a, ApJ, 708, 1002 Friesen, R. K., Di Francesco, J., Myers, P. C., et al. 2010b, ApJ, 718, 666 Galli, D., Walmsley, M., and Gonçalves, J. 2002, A&A, 394, 275 Genzel, R., and Stutzki, J. 1989, ARA&A, 27, 41 Goicoechea, J. R., Pety, J., Gerin, M., et al. 2009, A&A, 498, 771 Goldsmith, P. F., Bergin, E. A., and Lis, D. C. 1997, ApJ, 491, 615 Goodwin, S. P., Kroupa, P., Goodman, A., and Burkert, A. 2007, in Protostars and Planets V, ed. B. Reipurth, D. Jewitt, & K. Keil (Tuscon, AZ: Univ. Arizona Press), p. 133 Gordy, W., and Cook, R. L. 1984, Microwave Molecular Spectroscopy (New York: Wiley) Gredel, R., Reipurth, B., and Heathcote, S. 1992, A&A, 266, 439 Güsten, R., Nyman, L. Å., Schilke, P., et al. 2006, A&A, 454, L13 Harju, J., Haikala, L. K., Lehtinen, K., et al. 2006, A&A, 454, L55 Harju, J., Juvela, M., Schlemmer, S., et al. 2008, A&A, 482, 535 Hatchell, J., Thompson, M. A., Millar, T. J., and MacDonald, G. H. 1998, A&AS, 133, 29 Inutsuka, S.-I., and Miyama, S. M. 1992, ApJ, 388, 392 Johnstone, D., Rosolowsky, E., Tafalla, M., and Kirk, H. 2010, ApJ, 711, 655 Jørgensen, J. K., Schöier, F. L., and van Dishoeck, E. F. 2004, A&A, 416, 603 Juvela, M., Ristorcelli, I., Pelkonen, V.-M., et al. 2011, A&A, 527, A111 Klein, B., Philipp, S. D., Krämer, I., et al. 2006, A&A, 454, L29 Kovács, A. 2008, Proc. SPIE, 7020, 45 Kwon, W., Looney, L. W., Mundy, L. G., et al. 2009, ApJ, 696, 841 20 Lacy, J. H., Knacke, R., Geballe, T. R., and Tokunaga, A. T. 1994, ApJ, 428, L69 Ladd, E. F., Fuller, G. A., and Deane, J. R. 1998, ApJ, 495, 871 Lee, J.-E., Evans, N. J., II, Shirley, Y. L., and Tatematsu, K. 2003, ApJ, 583, 789 Loinard, L., Castets, A., Ceccarelli, C., et al. 2002, Planet. Space Sci., 50, 1205 Maret, S., Hily-Blant, P., Pety, J., et al. 2011, A&A, 526, A47 McKee, C. F. 1989, ApJ, 345, 782 Menten, K. M., Reid, M. J., Forbrich, J., and Brunthaler, A. 2007, A&A, 474, 515 Miettinen, O., Harju, J., Haikala, L. K., et al. 2009, A&A, 500, 845 (Paper I) Miettinen, O., Harju, J., Haikala, L. K., and Juvela, M. 2010, A&A, 524, A91 (Paper II) Miyake, K., and Nakagawa, Y. 1993, Icarus, 106, 20 Motte, F., and André, P. 2001, A&A, 365, 440 Müller, H. S. P., Schlöder, F., Stutzki, J., and Winnewisser, G. 2005, Journal of Molecular Structure, 742, 215 Ormel, C. W., Min, M., Tielens, A. G. G. M., et al. 2011, A&A, 532, A43 Ossenkopf, V., and Henning, Th. 1994, A&A, 291, 943 (OH94) Ostriker, J. 1964, ApJ, 140, 1056 Pagani, L., Vastel, C., Hugo, E., et al. 2009a, A&A, 494, 623 Pagani, L., Daniel, F., and Dubernet, M.-L. 2009b, A&A, 494, 719 Pickett, H. M., Poynter, I. R. L., Cohen, E. A., et al. 1998, J. Quant. Spec. Radiat. Transf., 60, 883 Planck Collaboration, Ade, P. A. R., Aghanim, N., et al. 2011, A&A, 536, A22 Rathborne, J. M., Simon, R., and Jackson, J. M. 2007, ApJ, 662, 1082 Rawlings, J. M. C., Taylor, S. D., and Williams, D. A. 2000, MNRAS, 313, 461 Reichertz, L. A., Weferling, B., Esch, W., and Kreysa, E. 2001, A&A, 379, 735 Remijan, A. J., Markwick-Kemper, A., and ALMA Working Group on Spectral Line Frequencies 2007, Bulletin of the American Astronomical Society, 38, 963 Roberts, H., and Millar, T. J. 2007, A&A, 471, 849 Savage, C., Apponi, A. J., Ziurys, L. M., and Wyckoff, S. 2002, ApJ, 578, 211 Schilke, P., Comito, C., Thorwirth, S., et al. 2006, A&A, 454, L41 Schmid-Burgk, J., Muders, D., Müller, H. S. P., and Brupbacher-Gatehouse, B. 2004, A&A, 419, 949 Schnee, S., and Goodman, A. 2005, ApJ, 624, 254 Schnee, S., Kauffmann, J., Goodman, A., and Bertoldi, F. 2007a, ApJ, 657, 838 Schnee, S., Caselli, P., Goodman, A., et al. 2007b, ApJ, 671, 1839 Schnee, S., Enoch, M., Noriega-Crespo, A., et al. 2010a, ApJ, 708, 127 Schnee, S., Enoch, M., Johnstone, D., et al. 2010b, ApJ, 718, 306 Schöier, F. L., van der Tak, F. F. S., van Dishoeck, E. F., and Black, J. H. 2005, A&A, 432, 369 Shetty, R., Kauffmann, J., Schnee, S., et al. 2009, ApJ, 696, 2234 Shirley, Y. L., Nordhaus, M. K., Grcevich, J. M., et al. 2005, ApJ, 632, 982 Shirley, Y. L., Huard, T. L., Pontoppidan, K. M., et al. 2011, ApJ, 728, 143 Siringo, G., Kreysa, E., De Breuck, C., et al. 2010, The Messenger, 139, 20 Tohline, J. E. 2002, ARA&A, 40, 349 van der Tak, F. F. S., Black, J. H., Schöier, F. L., et al. 2007, A&A, 468, 627 Vassilev, V., Meledin, D., Lapkin, I., et al. 2008a, A&A, 490, 1157 Vassilev, V., Henke, D., Lapkin, I., et al. 2008b, IEEE Microwave and Wireless Components Letters, pp. 55-60, Vol. 18, Number 1 Visser, A. E., Richer, J. S., Chandler, C. J., and Padman, R. 1998, MNRAS, 301, 585 Viti, S., Natarajan, S., and Williams, D. A. 2002, MNRAS, 336, 797 Watson, W. D., Anicich, V. G., and Huntress, W. T., Jr. 1976, ApJ, 205, L165 Williams, J. P., de Geus, E. J., and Blitz, L. 1994, ApJ, 428, 693 Williams, J. P., Bergin, E. A., Caselli, P., et al. 1998, ApJ, 503, 689 Wilson, T. L., and Rood, R. 1994, ARA&A, 32, 191 Woodall, J., Agúndez, M., Markwick-Kemper, A. J., and Millar, T. J. 2007, A&A, 466, 1197