content
stringlengths
0
113k
doi
stringlengths
14
40
pagenum
stringlengths
9
9
Astronomy &Astrophysics A&A 658, A135 (2022) https://doi.org/10.1051/0004-6361/202141662 © J. M. Winters et al. 2022 Molecules, shocks, and disk in the axi-symmetric wind of the MS-type AGB star RS Cancri? J. M. Winters1 , D. T. Hoai2 , K. T. Wong1 , W.-J. Kim3,4 , P. T. Nhung2 , P. Tuan-Anh2 , P. Lesaffre5 , P. Darriulat2, and T. Le Bertre6 1Institut de Radioastronomie Millimétrique (IRAM), 300 rue de la Piscine, Domaine Universitaire, 38406 St. Martin d’Hères, France e-mail: winters@iram.fr 2Department of Astrophysics, Vietnam National Space Center (VNSC), Vietnam Academy of Science and Technology (VAST), 18 Hoang Quoc Viet, Cau Giay, Ha Noi, Vietnam 3Instituto de Radioastronomía Milimétrica (IRAM), Av. Divina Pastora 7, Núcleo Central, 18012, Granada, Spain 4I. Physikalisches Institut, Universität zu Köln, Zülpicher Str. 77, 50937 Köln, Germany 5Laboratoire de Physique de l’École Normale Supérieure, 24 rue Lhomond, 75231 Paris, France 6LERMA, UMR 8112, CNRS and Observatoire de Paris, PSL Research University, 61 av. de l’Observatoire, 75014 Paris, France Received 29 June 2021 / Accepted 29 November 2021 ABSTRACT Context. The latest evolutionary phases of low- and intermediate-mass stars are characterized by complex physical processes like turbulence, convection, stellar pulsations, magnetic fields, condensation of solid particles, and the formation of massive outflows that inject freshly produced heavy elements and dust particles into the interstellar medium. Aims. By investigating individual objects in detail, we wish to analyze and disentangle the effects of the interrelated physical processes on the structure of the wind-forming regions around them. Methods. We use the Northern Extended Millimeter Array to obtain spatially and spectrally resolved observations of the semi- regular asymptotic giant branch (AGB) star RS Cancri and apply detailed 3D reconstruction modeling and local thermodynamic equilibrium radiative transfer calculations in order to shed light on the morpho-kinematic structure of its inner, wind-forming environment. Results. We detect 32 lines of 13 molecules and isotopologs (CO, SiO, SO, SO 2, H2O, HCN, PN), including several transitions from vibrationally excited states. HCN, H13CN, and millimeter vibrationally excited H 2O, SO,34SO, SO 2, and PN are detected for the first time in RS Cnc. Evidence for rotation is seen in HCN, SO, SO 2, and SiO(v=1). From CO and SiO channel maps, we find an inner, equatorial density enhancement, and a bipolar outflow structure with a mass-loss rate of 1107M yr1for the equatorial region and of2107M yr1for the polar outflows. The12CO/13CO ratio is measured to be 20on average, 242in the polar outflows and 193in the equatorial region. We do not find direct evidence of a companion that might explain this kind of kinematic structure, and explore the possibility that a magnetic field might be the cause of it. The innermost molecular gas is influenced by stellar pulsation and possibly by convective cells that leave their imprint on broad wings of certain molecular lines, such as SiO and SO. Conclusions. RS Cnc is one of the few nearby, low-mass-loss-rate, oxygen-rich AGB stars with a wind displaying both an equatorial disk and bipolar outflows. Its orientation with respect to the line of sight is particularly favorable for a reliable study of its morpho- kinematics. Nevertheless, the mechanism causing early spherical symmetry breaking remains uncertain, calling for additional high spatial- and spectral-resolution observations of the emission of different molecules in different transitions, along with more thorough investigation of the coupling among the different physical processes at play. Key words. stars: AGB and post-AGB – circumstellar matter – stars: mass-loss – stars: winds, outflows – stars: individual: RS Cnc – radio lines: stars 1. Introduction Mass-loss in red giants is due to a combination of stellar pulsations and radiation pressure on dust forming in dense shocked regions in the outer stellar atmosphere (e.g., Höfner & Olofsson 2018). Even if the basic principles are understood, a fully consistent picture – including the role of convection, the time-dependent chemistry, and a consistent description of dust formation – still needs to be developed. In particular, the contri- bution of transparent grains to the acceleration of matter close ?NOEMA data (FITS format) are only available at the CDS via anony- mous ftp to cdsarc.u-strasbg.fr (130.79.128.5 ) or via http: //cdsarc.u-strasbg.fr/viz-bin/cat/J/A+A/658/A135to the stellar photosphere (Norris et al. 2012) still needs to be assessed. The mechanisms shaping circumstellar environments around asymptotic giant branch (AGB) stars are vividly debated. Among them, magnetic fields (Matt et al. 2000; Duthu et al. 2017), bina- rity (Theuns & Jorissen 1993; Mastrodemos & Morris 1999; Decin et al. 2020), stellar rotation (Dorfi & Höfner 1996), and common-envelope evolution (Olofsson et al. 2015; Glanz & Perets 2018) have been considered. A major difficulty is to explain the observed velocity field in axi-symmetrical sources, with larger velocities at high lati- tudes than at low latitudes (Hoai et al. 2014; Nhung et al. 2015b). Also, recent observations of rotating structures and streams bring additional conundrums (Tuan-Anh et al. 2019; Hoai et al. 2019). A135, page 1 of 27 Open Access article, published by EDP Sciences, under the terms of the Creative Commons Attribution License ( https://creativecommons.org/licenses/by/4.0 ), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
10.1051_0004-6361_202141662
page_0000
A&A 658, A135 (2022) We have concentrated our efforts on two relatively close (d150pc) sources that show composite profiles in CO rota- tional lines (Winters et al. 2003): EP Aqr (Winters et al. 2007, hereafter referred to as W2007), and RS Cnc (Libert et al. 2010). Data obtained at IRAM show that these two sources have an axi-symmetrical structure with a low-velocity ( 2 km s1) wind close to the equatorial plane, and faster ( 8 km s1) outflows around the polar axes (Hoai et al. 2014; Nhung et al. 2015b). For EP Aqr, W2007 find a radial dependence of the density show- ing intermediate maxima. Additional data obtained with ALMA (Nhung et al. 2019b; Homan et al. 2018b) reveal a spiral structure explaining the earlier W2007 results. RS Cnc is one of the best examples of the interaction between the stellar wind from an AGB star and the surrounding interstel- lar medium (Hoai et al. 2014). Its high declination makes RS Cnc an ideal target for the Northern Extended Millimeter Array (NOEMA). Previous studies based on IRAM data show that it is a twin of EP Aqr, but observed at a different angle, with a polar axis inclined at about 30with respect to the line of sight (Libert et al. 2010; Hoai et al. 2014; Nhung et al. 2015b). This is favor- able for studying polar and equatorial structures simultaneously, whereas the different viewing angle between EP Aqr and RS Cnc can be exploited to discriminate between different models in explaining the observed composite CO line profiles (Le Bertre et al. 2016). In contrast to EP Aqr, technetium is detected in the atmosphere of RS Cnc (Lebzelter & Hron 1999), proving that it is evolving along the thermal pulsing asymptotic giant branch (TP-AGB) in the Hertzsprung-Russell (HR) diagram. From a chemical point of view, RS Cnc is in a slightly more advanced evolutionary stage on the AGB, as indicated by its spectral clas- sification as an MS star (see below) and by a higher photospheric ratio of12C/13C (35; Smith & Lambert (1986), but see Sect. 4.1 for an improved evaluation based on CO rotational lines from the circumstellar environment). RS Cnc is a semi-regular variable star with periods of 122 d and248 days (Adelman & Dennis 2005), located at a distance of150pc (Gaia Collaboration 2021; Bailer-Jones et al. 2021). It is listed as S-star CSS 589 in Stephenson (1984) based on its spectral classification M6S given in Keenan (1954). With its weak ZrO bands, its chemical type is intermediate between M and S (Keenan 1954). The stellar temperature is estimated toT3200 K and its luminosity is L4950 L (Dumm & Schild 1998). From CO rotational line observations, two circum- stellar wind components were identified: an equatorial structure expanding at about 2 km s1and a bipolar outflow reaching a terminal velocity of vexp8km s1(Libert et al. 2010; Hoai et al. 2014), carrying mass-loss rates of 4108M yr1and 8108M yr1, respectively (see Sect. 4.1 for an improved value of the mass-loss rate derived here). Lines of12CO, 13CO, SiO, and HI were detected from previous observations at millimeter (mm) and radio wavelengths (Nyman et al. 1992; Danilovich et al. 2015; de Vicente et al. 2016; Gérard & Le Bertre 2003; Matthews & Reid 2007). NOEMA was recently equipped with the wide band cor- relator PolyFiX, covering a total bandwidth of 15.6 GHz and therefore offering the potential to observe several lines from different species simultaneously. In this paper we present new data obtained with NOEMA in D- and A-configuration, com- plemented by short spacing observations obtained at the IRAM 30m telescope. Observational details are summarized in Sect. 2 and our results are presented in Sect. 3. Section 4 contains a discussion of the morphological structures and compares them to similar structures found in EP Aqr. Our conclusions are summarized in Sect. 5.2. Observations New observations of RS Cnc have been obtained in CO(2–1) with NOEMA/WideX in the (extended) nine-antenna A- configuration in December 2016 (Nhung et al. 2018) and with NOEMA/PolyFiX in the (compact) nine-antenna D- configuration during the science verification phase of PolyFiX in December 2017 and in the ten-antenna A-configuration in February 2020. The WideX correlator covered an instanta- neous bandwidth of 3.8 GHz in two orthogonal polarizations with a channel spacing of 2 MHz. Additionally, up to eight high-spectral resolution units could be placed on spectral lines, providing channel spacings down to 39 kHz. WideX was decom- missioned in September 2017 and replaced in December 2017 by the new correlator PolyFiX. This new correlator simultaneously covers 7.8 GHz in two sidebands and for both polarizations, and provides a channel spacing of 2 MHz throughout the 15.6 GHz total bandwidth. In addition, up to 128 high-spectral-resolution “chunks” can be placed in the 15.6 GHz-wide frequency range covered by PolyFiX for both polarizations, each providing a fixed channel spacing of 62.5 kHz over their 64 MHz bandwidth. RS Cnc was observed with two individual frequency setups covering a total frequency range of 32 GHz in the 1.3 mm atmo- spheric window (see Fig. 1). We used the two quasars J0923+282 and 0923+392 as phase and amplitude calibrators; these were observed every20 min. Pointing and focus of the telescopes was checked about every hour, and corrected when necessary. The bandpass was calibrated on the strong quasars 3C84 and 3C273, and the absolute flux scale was fixed on MWC349 and LkHa101, respectively. The accuracy of the absolute flux calibration at 1.3 mm is estimated to be better than 20%. In order to add the short spacing information filtered out by the interferometer, in May and July 2020 we observed at the IRAM 30m telescope maps of 10by 10using the On-The-Fly (OTF) mode. This turned out to be necessary for the12CO(2– 1) and13CO(2–1) lines but was not needed for the SiO lines, whose emitting region was found to be smaller than 300. In the case of the12CO(2–1) and13CO(2–1) lines, the interferometer filters out large-scale structures that account for about two-thirds and three-quarters, respectively, of the total line flux, informa- tion that is recovered by adding the short spacing data from the OTF map. A comparison of the respective line profiles is shown in Fig. A.1. The data were calibrated and imaged within the GILDAS1 suite of software packages using CLIC for the NOEMA data calibration and the uvtable creation, CLASS for calibrating the OTF maps, and the MAPPING package for merging and subse- quent uvfitting, imaging, and self-calibration of the combined data sets. Continuum data were extracted for each sideband of the two frequency setups individually by filtering out spectral lines, and then averaging over 400 MHz bins to properly rescale theuvcoordinates to the mean frequency of each bin. Phase self-calibration was performed on the corresponding continuum data. The gain table containing the self-calibration solutions was then applied to the spectral line uvtables using the SELFCAL procedures provided in MAPPING. The resulting data sets were imaged applying either natu- ral weighting, or, on the high-signal-to-noise (S/N) cubes, by applying robust weighting with a threshold of 0.1 to increase the spatial resolution by typically a factor 2. The resulting dirty maps were then CLEANed using the Hogbom algorithm (Högbom 1974). 1https://www.iram.fr/IRAMFR/GILDAS A135, page 2 of 27
10.1051_0004-6361_202141662
page_0001
J. M. Winters et al.: Molecules, shocks, and disk in the axi-symmetric wind of the MS-type AGB star RS Cancri Table 1. Properties of the combined data sets for all detected lines. Line Frequency Eu=k Peak flux FWHP beam size PA 1 noise vel.res Comments(a) (GHz) (K) (Jy) (arcsec) (arcsec2) (deg) (mJy beam1) ( km s1) 12CO(2–1) 230.538000 16.6 53.971 10.841 6.160.01 0.480.30 36 2.88 0.5 A+D+30m, rw 13CO(2–1) 220.398684 15.9 4.693 0.948 7.200.01 0.500.31 35 2.79 0.5 A+D+30m, rw SiO(v=0,5–4) 217.104919 31.3 17.464 3.523 1.710.01 0.510.32 36 3.38 0.5 A+D, rw, sc SiO(v=1,5–4) 215.596018 1800.2 0.105 0.025 0.190.02 0.580.43 38 1.71 1.0 A, nw, sc, Feb 2020: no maser SiO(v=1,5–4) 215.596018 1800.2 0.105 0.025 0.190.02 2.101.80 0 2.71 0.5 D, nw, sc, Dec 2017: maser SiO(v=2,5–4) 214.088575 3552.1 0.013 0.005 0.350.09 1.000.74 35 1.03 3.0 A+D, nw, double peak profile (?) SiO(v=0,6–5) 260.518009 43.8 23.906 4.817 1.620.01 0.430.26 32 3.39 0.5 A+D, rw, sc SiO(v=1,6–5) 258.707324 1812.7 0.168 0.038 0.110.01 0.600.42 26 1.96 1.0 A+D, nw, sc 29SiO(v=0,5–4) 214.385752 30.9 5.372 1.083 1.190.01 0.520.32 37 1.12 3.0 A+D, rw, sc Si17O(v=0,6–5) 250.744695 42.1 0.340 0.076 0.880.04 1.901.50 36 4.15(b)3.0 D, nw, sc tentative identification 29Si17O(v=0,6–5) 247.481525 41.6 0.020 0.008 0.730.44 1.901.50 26 2.10 3.0 D, nw, sc, tentative detection SO(5(5)–4(4)) 215.220653 44.1 0.455 0.093 0.790.01 0.510.32 36 1.16 3.0 A+D, rw, sc SO(6(5)–5(4)) 219.949442 35.0 0.634 0.130 0.800.01 0.500.31 36 1.17 3.0 A+D, rw, sc SO(6(6)–5(5)) 258.255826 56.5 0.870 0.178 0.740.01 0.430.27 32 1.59 3.0 A+D, rw, sc SO(7(6)–6(5)) 261.843721 47.6 1.168 0.238 0.780.01 0.430.26 32 1.38 3.0 A+D, rw, sc 34SO(6(5)–5(4)) 215.839920 34.4 0.030 0.009 0.920.11 0.910.80 69 0.86 3.0 A+D, nw 34SO(5(6)–4(5)) 246.663470 49.9 0.026 0.009 0.930.14 0.690.48 26 0.99 3.0 A+D, nw SO2(16(3,13)–16(2,14)) 214.689394 147.8 0.021 0.006 0.500.06 0.900.68 37 1.06 3.0 A+D, nw, sc SO2(22(2,20)–22(1,21)) 216.643304 248.4 0.023 0.007 0.380.05 0.890.67 36 1.11 3.0 A+D, nw, sc SO2(28(3,25)–28(2,26)) 234.187057 403.0 0.022 0.006 0.190.05 0.710.56 46 1.28 3.0 A+D, nw, sc SO2(14(0,14)–13(1,13)) 244.254218 93.9 0.043 0.011 0.430.03 0.690.49 27 1.00 3.0 A+D, nw, sc SO2(10(3, 7)–10(2, 8)) 245.563422 72.7 0.025 0.007 0.360.04 0.690.49 28 1.00 3.0 A+D, nw, sc SO2(15(2,14)–15(1,15)) 248.057402 119.3 0.015 0.005 0.260.06 0.690.48 27 1.07 3.0 A+D, nw, sc SO2(32(4,28)–32(3,29)) 258.388716 531.1 0.020 0.006 0.210.04 0.630.45 25 1.10 3.0 A+D, nw, sc SO2( 9(3, 7)– 9(2, 8)) 258.942199 63.5 0.026 0.008 0.490.05 0.640.45 26 1.08 3.0 A+D, nw, sc SO2(30(4,26)–30(3,27)) 259.599448 471.5 0.022 0.005 0.120.03 0.630.45 25 1.03 3.0 A+D, nw, sc SO2(30(3,27)–30(2,28)) 263.543953 459.0 0.019 0.006 0.160.04 0.610.42 26 1.25 3.0 A+D, nw, sc SO2(34(4,30)–34(3,31)) 265.481972 594.7 0.020 0.006 0.190.04 0.610.42 26 1.27 3.0 A+D, nw, sc H2O(v2=1,5(5,0)–6(4,3)) 232.686700(c)3462.0 0.0290.007 unresolved 0.71 0.57 47 1.17 3.0 A+D, nw, sc, JPL H2O(v2=1,7(7,0)–8(6,3)) 263.451357(d)4474.7 0.0210.005 unresolved 0.61 0.42 26 1.17 3.0 A+D, nw, sc, JPL HCN(3–2) 265.886434 25.5 1.116 0.234 0.760.01 0.420.26 32 4.80 0.5 A+D, rw, sc H13CN(3–2) 259.011798 24.9 0.041 0.011 0.710.05 0.640.45 26 0.95 3.0 A+D, nw, sc PN(N=5–4,J=6–5) 234.935694 33.8 0.028 0.009 0.800.10 0.700.56 47 1.00 3.0 A+D, nw, sc Notes. Line frequencies and upper level energies are from the CDMS (Müller et al. 2005), unless otherwise stated. The quoted flux uncertainties include the rms of the fits and the absolute flux calibration accuracy of 20%, the uncertainties quoted for the source sizes refer to the rms errors of the Gaussian fits (see text).(a)A: NOEMA A-configuration, D: NOEMA D-configuration, 30 m: short spacing data, rw: robust weighting, nw: natural weighting, sc: self-calibrated, JPL: Spectral line catalog by NASA/JPL (Pickett et al. 1998).(b)Increased noise at band edge.(c)Belov et al. (1987).(d)Pearson et al. (1991). The beam characteristics and sensitivities of the individual combined data sets from A- and D-configuration (and including the pseudo-visibilities from the OTF maps, where appropriate) are listed in Table 1 for all detected lines. 3. Results The PolyFiX data, covering the frequency ranges 213–221 GHz (setup1, LSB), 228-236 GHz (setup1, USB), 243–251 GHz (setup2, LSB), and 258–266 GHz (setup2, USB) with two setups (see Fig. 1), showed different lines of CO and SiO, and, for the first time, many lines of species like SO, SO 2, HCN, and PN and some of their isotopologs. Furthermore, the data confirmed the H2O line at 232.687 GHz already detected serendipitously withWideX in 2016, with a second H 2O line at 263.451 GHz seen for the first time in RS Cnc. All lines covered by the same setup (1 or 2, see Fig. 1) share the same phase-, amplitude-, and flux calibration. All 32 detected lines are listed in Table 1. 3.1. Continuum Figure 2 shows the continuum map from A-configuration only, using robust weighting to increase the spatial resolution to 0:39000:2200at PA 28. After self-calibration, S/N = 492 is obtained. The continuum source is unresolved, a point source fit results in a flux at 247 GHz of 23.65 4.7 mJy (where the quoted error accounts for the accuracy of the absolute flux cali- bration of 20%) and a source position at RA = 09:10:38.780 and A135, page 3 of 27
10.1051_0004-6361_202141662
page_0002
A&A 658, A135 (2022) Fig. 1. Overview of the frequency ranges observed with PolyFiX using two spectral setups (setup1: red and setup2: blue, respectively). Lower diagrams : zoom onto the individual spectra covering 7.8 GHz each. Upper row : setup1, lower row : setup2. The central 20 MHz at the border between inner and outer baseband are blanked out, i.e., set to zero, as this region is contaminated by the LO2 separation of the 8 GHz-wide IF in the IF processor (“LO2 zone”). Dec = 30:57:46.62 in February 2020. All line data cubes dis- cussed in the remainder of this paper are re-centered on this continuum position. The source position is offset from the J2000 coordinates by 0.2600in RA and by0.6800in Dec, consistent with the proper motion of RS Cnc ( 10.72 mas yr1in RA and33.82 mas yr1 in Dec, Gaia Collaboration 2021; Bailer-Jones et al. 2021). From the PolyFiX data, spanning a total frequency range of about 53 GHz, we determine a spectral index of 1.99 0.09 for RS Cnc in the 1 mm range, which is fully consistent with a black body spectrum of the continuum (see also Libert et al. 2010). 3.2. Detected molecules and lines Within the total frequency coverage of about 32 GHz, we detect 32 lines of 13 molecules and isotopologs, including several tran- sitions from vibrationally excited states. All these lines are listed in Table 1 and are presented in the following sections. The peak flux and FWHP of the line-emitting regions, as listed in Table 1,are determined by circular Gaussian fits in the uv-plane to the central channel (if the source is (partially) spatially resolved) or by point-source fits to the central channel (if the source is unre- solved). All line profiles shown in the following sections in Fig. 3 and Figs. 5 through 14 are integrated over square apertures whose sizes are given in each figure caption. Two-component profiles are seen in CO and13CO only, and not in any other of the lines detected here. We looked for but did not detect the vibrationally excited 12CO(v=1, 2–1) line, nor do we detect C18O(2–1), result- ing in 3upper limits for the line peaks of 6 mJy beam1and 3 mJy beam1, respectively (the12CO(v=1, 2–1) line was not covered in our A-configuration data). 3.2.1. CO The profiles of12CO(2–1),13CO(2–1) (see Fig. 3), and12CO(1– 0) (see Libert et al. 2010) show a very distinct shape composed of a broad component that extends out to vlsr;8 km s1and A135, page 4 of 27
10.1051_0004-6361_202141662
page_0003
J. M. Winters et al.: Molecules, shocks, and disk in the axi-symmetric wind of the MS-type AGB star RS Cancri Fig. 2. Continuum map around 247 GHz from A-configuration. Con- tours are plotted in 100 steps, where 1 is 47.6 Jy beam1. The synthesized beam is indicated in the lower left corner. Fig. 3. CO line profiles, showing a two-component structure. Left: 12CO(2–1). Right :13CO(2–1). A-configuration and D-configuration are merged, OTF data are added, and the spectral resolution is 0.5 km s1. The CO emission is integrated over the central 22002200, i.e., over the full field of view of the NOEMA antennas at 230 GHz. Fig. 4. Sketch of the geometrical structure of the wind components as inferred from the current data (see Sect. 4.1). The sketch is not to scale: there is a smooth transition between the equatorial enhancement and the polar outflows. a narrow component indicating velocities of 2 km s1with respect tovlsr;=7km s1. Velocity-integrated intensity maps of CO are shown in Fig. 18, indicating a clear kinematic structure in the north–south direction. In Fig. 4, we present a schematic representation of the geometrical structure of RS Cnc as implied by the data; see Sect. 4.1. The CO emitting region is spatially extended, consisting of a dense equatorial structure that corre- sponds to the low-velocity expansion and an inclined, bipolar Fig. 5. Profiles of SiO ground-state and first vibrationally excited state lines. Left: SiO(6–5): upper :v=1,lower :v=0. A-configuration and D- configuration merged. Right : SiO(5–4): upper :v=1, D-configuration (black) and A-configuration (red), lower :v=0, A-configuration and D- configuration merged. The spectral resolution is 1 km s1for (v=1) and 0.5 km s1for the (v=0) lines, respectively. The emission is integrated over the central 500500aperture. structure corresponding to an outflow at a projected velocity of 8 km s1. These structures were discussed in Hoai et al. (2014) based on Plateau de Bure data obtained on12CO(2–1) and12CO(1–0) that had a spatial resolution of about 100. The model built by these latter authors was later refined by Nhung et al. (2018) based on12CO(2–1) data obtained with the WideX correlator in NOEMA’s A-configuration, providing a spatial res- olution of 0:44000:2800. Nhung et al. (2018) find a position angle of the projected bipolar outflow axis of !=7(measured counter-clockwise from north) and an inclination angle of the outflow axis with respect to the line of sight of i=30. The CO distribution is further investigated in Sect. 4.1 below. Such a structure had already been found in the S-type star 1 Gru (Sahai 1992), which was later confirmed by higher spatial resolution observations using ALMA (Doan et al. 2017). This object has a G0V companion (Feast 1953) and possibly a sec- ond, much closer companion (Homan et al. 2020). In Hoai et al. (2014), we reported for RS Cnc the possible presence of a com- panion seen in the12CO(1–0) channel maps at velocities around 6.6 km s1and located about 100west-northwest of the contin- uum source. The new data allow for a more detailed study of this feature, which is presented in Sect. 4.1. 3.2.2. SiO We detect a suite of28Si16O (henceforth SiO) transitions, includ- ing the vibrational ground-state lines of SiO(5–4) and SiO(6–5), the first and second vibrationally excited state of SiO(5–4), and the first vibrationally excited state of SiO(6–5). All SiO pro- files are shown in Figs. 5 and 6. The spatial region emitting the vibrational ground-state lines extends out to about 200from the continuum peak (see Table 1, Fig. 18, and Sect. 4.2). Interest- ingly, we detect a strong maser component on the SiO( v=1, J=5–4) line at vlsr14km s1in the data obtained in December 2017, which had completely disappeared when we re-observed RS Cnc in February 2020 (see Fig. 5). Such behav- ior is well known for pulsating AGB stars, and lends support to A135, page 5 of 27
10.1051_0004-6361_202141662
page_0004
A&A 658, A135 (2022) Fig. 6. Profile around the SiO( v=2,J=5–4) line frequency. A- configuration and D-configuration merged. The spectral resolution is 1 km s1and the emission is integrated over the central 100100aperture. the idea that the SiO masers are excited by infrared pumping as opposed to collisional pumping (see, e.g., Pardo et al. 2004). The SiO(v=2,J=5–4) line is detected above the 3 level of 3 mJy beam1over a broad range of Doppler velocities from at least5 to 18 km s1(Fig. 6). Given its high excitation energy (3500 K), we expect this line to trace exclusively the inner- most region around RS Cnc, as was the case in oCet, where SiO(v=2) absorption and emission was spatially resolved by ALMA (Wong et al. 2016). Its broad line width suggests that it may trace the same high-velocity wings seen in other detected SiO lines (Sect. 3.3). However, our detection is too weak to allow for a detailed study of the morpho-kinematics of the emission. At the upper edge of the LSB of setup 2 at 250.744 GHz, we serendipitously detect a strong line that we identify as ground-state Si17O(6–5) at 250.7446954 GHz (Müller et al. 2013) from the Cologne Database of Molecular Spectroscopy (CDMS2, Müller et al. 2005); the profile is shown in Fig. 7. This line and other transitions of Si17O have already been detected in a number of well-studied objects, such as the S-type star W Aql (De Beck & Olofsson 2020), the M-type star R Dor (De Beck & Olofsson 2018), and the evolved, high-mass-loss- rate oxygen-rich star IK Tau (Velilla Prieto et al. 2017). No other Si17O transitions are covered in our setups, but there is a highly excited H 2O line at 250.7517934 GHz ( v2=2,J(Ka;Kc)= 9(2,8)–8(3,5); Eu=k=6141 K) listed in the JPL catalog3and pre- dicted by Yu et al. (2012) from the Bending-Rotation approach analysis. If the detected line was H 2O emission, it would be redshifted from the systemic velocity by about 9 km s1. As indi- cated by the modeling of Gray et al. (2016), the 250.752 GHz line may exhibit strong maser action in regions of hot gas (Tkin=1500 K) with cool dust ( Td1000 K). While we can- not unequivocally exclude some contamination from a potential new, redshifted H 2O maser, we consider Si17O a more likely identification of the 250.744 GHz emission. From the respective integrated line intensities of Si16O(6–5) and Si17O(6–5), which are163 Jy km s1and3 Jy km s1, and taking the difference of the Einstein coefficients of the transitions into account, we estimate the isotopolog ratio16O/17O50, assuming equal exci- tation conditions for both transitions and optically thin emission of both lines. This value is much lower than the solar isotopic ratio of2700 (Lodders et al. 2009) due to dredge-up events (Karakas & Lattanzio 2014; Hinkle et al. 2016) and is broadly consistent with those obtained in the M-type star R Dor and the S-type star W Aql (61–74; De Beck & Olofsson 2018, 2020). The initial mass of RS Cnc is about 1:5M (Libert et al. 2010)4, 2https://cdms.astro.uni-koeln.de 3https://spec.jpl.nasa.gov/ftp/pub/catalog/catform. html 4As quoted in Libert et al. (2010), the value of 1:5M was esti- mated by Busso & Palmerini (their priv. comm.) using the FRANEC ).Fig. 7. Line profiles of SiO isotopologs. Upper left : profile of the 247.482 GHz line, possibly29Si17O(6–5); D-configuration, only. Lower left: Si17O(6–5): D-configuration, only (line was not covered in A- configuration). Right :29SiO(5–4); A-configuration and D-configuration merged. The spectral resolution in all cases is 3 km s1and the emission is integrated over the central 500500aperture. which is in the same range as R Dor ( 1:4M ; De Beck & Olofsson 2018) and W Aql ( 1:6M ; De Nutte et al. 2017) that gives a16O/17O ratio of<1000 (Hinkle et al. 2016). However, we note that the oxygen isotopic ratio (16O/17O) derived from the line intensity ratio is likely underestimated if the Si16O line is not optically thin, as has been shown in De Beck & Olofsson (2018), who obtained a value of 400in R Dor with radiative transfer modeling. Indeed, we demonstrate in Sect. 4.2 that the Si16O emission in RS Cnc is optically thick, especially within a projected radius of 100. A photospheric16O/17O ratio of 710 in RS Cnc (=HR 3639) was estimated by Smith & Lambert (1990) from the spectra of near-infrared overtone band transitions of C16O and C17O, which is probably a more realistic ratio. We do not cover C17O(2–1) in our setups and therefore cannot give an independent estimate of the16O/17O ratio. As Si18O(6–5) and C18O(2–1) are either not covered or not detected, there is not enough information from our data to obtain a meaningful con- straint on the initial stellar mass from oxygen isotopic ratios (e.g. from the17O/18O ratio; De Nutte et al. 2017). We detect a line at 247.482 GHz at low S/N that might be identified as29Si17O(v=0,J=6–5) at 247.4815250 GHz based on the line list by Müller et al. (2013) and used in the CDMS (see Fig. 7). However, in contrast to Si17O(6–5),29Si17O(6–5) has never been detected; only higher-J lines of29Si17O have been tentatively detected in R Dor ( J=7–6 and J=8–7, De Beck & Olofsson 2018). More specifically, the 247.482 GHz line is seen with an integrated line intensity of 0:08Jy km s1in our D-configuration data only, observed in December 2017, but it does not show up in the A-configuration data, taken in February 2020. This may largely be due to the much reduced brightness stellar evolution code (Cristallo et al. 2011) and the molecular abun- dances determined by Smith & Lambert. Smith & Lambert (1990) reported oxygen isotopic ratios of16O/17O=710 and16O/18O=440 in RS Cnc (their Table 9). The17O/18O ratio of 0.62 corresponds to an initial mass of 1.4–1.5 M in the comparative study of De Nutte et al. (2017), who investigated the17O/18O isotopic ratio as a sensitive function of initial mass of low-mass stars based on the models of Stancliffe et al. (2004), Karakas & Lattanzio (2014), and the FRANEC model. A135, page 6 of 27
10.1051_0004-6361_202141662
page_0005
J. M. Winters et al.: Molecules, shocks, and disk in the axi-symmetric wind of the MS-type AGB star RS Cancri ). Fig. 8. HCN line profiles. Left: HCN(3–2); A-configuration and D- configuration merged with a spectral resolution of 0.5 km s1.Right : H13CN(3–2); A-configuration and D-configuration merged with a spec- tral resolution of 3 km s1. The emission of both lines is integrated over the central 200200. sensitivity in the A-configuration, which is a factor of approxi- mately 15 smaller because of the smaller synthesized beam area rather than some variable maser action in this line. Based on the D-configuration data, the source position of the 247.482 GHz emission appears slightly offset toward the northwest direction from the Si17O(6–5) emission. Further data on29Si17O, possibly covering the J=6–5,J=7–6, and J=8–7 transitions, would be needed to draw any firm conclusion. 3.2.3. HCN We clearly detect the HCN(3–2) and H13CN(3–2) lines; the profiles are displayed in Fig. 8, and velocity-integrated inten- sity maps of both species are shown in Fig. B.1. Both lines are slightly spatially resolved and a circular Gaussian fit to HCN(3–2) gives a peak flux of 1.12 Jy and a FWHP size of 0.7600 on the merged data. To our knowledge, this is the first detection of HCN and H13CN in RS Cnc (see Sect. 4.4). From the first- moment map (shown in Fig. 17, left), a clear velocity pattern is evident that indicates possible rotation in the HCN-emitting region (see Sect. 3.4). Also, the velocity-integrated intensity maps presented in Fig. B.1 show a clear kinematic structure in the east–west direction. Formation of the HCN molecule in oxygen-rich environ- ments is further discussed in Sect. 4.4. A modeling using the 1D local thermodynamic equilibrium (LTE) radiative transfer code XCLASS (Möller et al. 2017, see Appendix D) gives a column density for HCN in RS Cnc of N HCN=1:61015cm2, cor- responding to an abundance of X(HCN/H 2)=6:6107. This value is well within the range found for other M- and S-type stars as modeled by Schöier et al. (2013), who find X(HCN/H 2) equal to a few times 107(for more details see Sect. 4.4 and Appendix D). 3.2.4. H 2O The WideX spectrum obtained in A-configuration in Decem- ber 2016 serendipitously revealed a line at 232.687 GHz that we ascribe to the J(Ka,Kc)=5(5,0)–6(4,3) transition of o-H 2O in thev2=1vibrational state. The H 2O source is weak and seems still unresolved within the synthesized beam of 0:5000:3400 obtained in the A-configuration in February 2020, consistent with its high upper-state energy of 3462 K. The line profile is shown in Fig. 9. With the follow-up observations employing PolyFiX in D-configuration and A-configuration we also cov- ered and detected the 263.451 GHz o-H 2Ov2=1,J(Ka,Kc)= 7(7,0)–8(6,3) line (Fig. 9, right; Eu=k=4475 K). Both lines are resampled to a resolution of 3 km s1, data are merged from A-configuration and D-configuration, and the emission is Fig. 9. H2O line profiles. Left: H 2O line at 232.687 GHz. Right : H 2O line at 263.451 GHz. Data are merged from A-configuration and D- configuration, the spectral resolution is 3 km s1, and the emission of both lines is integrated over the central 100100aperture. integrated over an aperture of 100100. Intensity maps of both lines are shown in Fig. B.2, testifying to the compactness of the H2O-emitting region. These are the first detections of millimeter vibrationally excited H 2O emission in RS Cnc. We note that the 22 GHz H2O maser in the ground state was tentatively detected by Szymczak & Engels (1995) in one of the two epochs they cov- ered, but the 22 GHz line is not detected in other observations (Dickinson et al. 1973; Lewis 1997; Han et al. 1995; Yoon et al. 2014). RS Cnc also shows clear photospheric H 2O absorption at 2:7m (Merrill & Stein 1976; Noguchi & Kobayashi 1993), and at1:3m (7500 cm1; Joyce et al. 1998), although the H 2O band near 900 nm is not detected (Spinrad et al. 1966). Both the 232 and 263 GHz water lines have upper levels belonging to the so-called transposed backbone in the v2=1 vibrationally excited state of H 2O, that is Ka=JandKc=0or 1 (see Fig. 1 of Alcolea & Menten 1993). The 232 GHz line was first detected in evolved stars together with the 96 GHz line from another transposed backbone upper level by Menten & Melnick (1989) toward the red supergiant VY CMa and the AGB star W Hya. The latter is an M-type star with a similar mass-loss rate to RS Cnc. The authors find that the 232 GHz line emission in both stars may be of (quasi-)thermal nature while the 96 GHz line clearly showed maser action. The (unpublished) detection of the 263 GHz line was mentioned in Alcolea & Menten (1993), who also described a mechanism that may lead to a system- atic overpopulation of the transposed backbone upper levels in thev2=1state of H 2O in the inner region of circumstellar envelopes. If the vibrational decay routes (to the ground state) of the transposed backbone upper levels become more optically thick than the lower levels in the v2=1state, then differential radiative trapping may cause population inversion of these lines. Additional vibrationally excited H 2O emission lines from trans- posed backbone upper levels were predicted and later detected in VY CMa by Menten et al. (2006) and Kami ´nski et al. (2013). We observed the 232 GHz line in RS Cnc at three epochs (December 2016, December 2017, and February 2020) and the 263 GHz line at the latter two epochs, and the emission appears to be stable in time for both lines. The profiles appear to be very similar, both are broad, even broader than the (ground-state) lines of other species reported here, and there is no sign for any narrow com- ponent in either of the two profiles at any of the epochs. As the lines should arise from a region very close to the star – compat- ible with their broad widths; see Sect. 3.3 – one might expect to see time variations due to the varying density and radiation field caused by the stellar pulsation, in particular if the emission were caused by maser action, as seen on the SiO( v=1;5–4) line observed in December 2017 (see Fig. 5). Also, the model- ing of Gray et al. (2016) shows only very little inversion of the A135, page 7 of 27
10.1051_0004-6361_202141662
page_0006
A&A 658, A135 (2022) Fig. 10. Profiles of the four detected SO lines, with A-configuration and D-configuration merged. The spectral resolution is 3 km s1and the emission is integrated over the central 200200aperture. Fig. 11. Profiles of the two34SO lines detected here, with A- configuration and D-configuration merged. The spectral resolution is 3 km s1and the emission is integrated over the central 200200aperture. involved level populations for the 263 GHz H 2O transition. We therefore think that both lines could be thermally excited. A def- inite assessment of the nature of the vibrationally excited H 2O emission would however require some detailed modeling of the emission, together with high-sensitivity monitoring of the line profiles with high spectral resolution, possibly including other H2O lines from transposed backbone upper levels and/or known maser lines for comparison, which is beyond the scope of the present paper. 3.2.5. SO Four lines of SO are detected (see Fig. 10) along with two lines of the isotopolog34SO (Fig. 11). These represent the first detections of SO and34SO in RS Cnc. SO has been observed in several M- type stars, including R Dor and W Hya, (Danilovich et al. 2016)), but remains undetected in S-type stars (e.g., W Aql, Decin et al. 2008; De Beck & Olofsson 2020). All SO lines detected here are slightly spatially resolved with a FWHP around 0:800and therefore seem to be emitted from the same region as HCN. Velocity-integrated intensity maps of SO are shown in Fig. B.3. The SO lines show the same velocity pattern (indicating rota- tion) as HCN, although the velocity resolution of the SO lines is only 3 km s1; see Fig. B.3 and the first-moment map in the right panel of Fig. 17. Using the integrated line strengths of SO(6(5)–5(4)) and34SO(6(5)–5(4)) found here ( 4.69 Jy km s1and 0.20 Jy km s1, respectively) and taking the difference of the Einstein coefficients of the transitions into account, weestimate the isotopolog ratio32SO/34SO23, assuming equal excitation conditions for both transitions and optically thin emission of both lines. This value is in good agreement with the values of 21.68:5and 18.55:8derived from the radiative transfer models for M-type stars by Danilovich et al. (2016, 2020), respectively. We note that, for the S-type star W Aql, an Si32S/Si34S isotopolog ratio of 10.6 2:6was derived by De Beck & Olofsson (2020). As32S is mainly produced by oxygen burning in massive stars and, to a lesser extent, in type Ia supernovae, and as34S is formed by subsequent neutron capture (e.g., Nomoto et al. 1984; Wilson & Matteucci 1992; Timmes et al. 1995; Hughes et al. 2008), the32S/34S isotopic ratio remains virtually unaltered during AGB evolution (see, e.g. tables in the FRUITY5database, Cristallo et al. 2011) and there- fore should reflect the chemical initial conditions of the natal cloud from which the star has formed. The spread in the isotopic ratio seen among the different AGB stars mentioned above would then rather be indicative of the Galactic environment in which the star has formed (see, e.g., Chin et al. 1996; Humire et al. 2020) instead of reflecting any evolutionary effect. For the low-mass-loss-rate M-type stars R Dor and W Hya, Danilovich et al. (2016) reproduce their observed line profiles best with centrally peaked SO (and SO 2) distributions, consistent with the maps presented in Fig. B.3. 3.2.6. SO 2 In SO 2, 11 lines are detected; their parameters are summarized in Table 2, and all profiles are shown in Fig. C.1. These are the first detections of SO 2in RS Cnc. A previous survey with the IRAM 30 m telescope by Omont et al. (1993) did not detect SO2in RS Cnc with an rms noise of 0.052 K (or 0:25Jy at 160.8 GHz). As an example, we show the SO 2(14(0,14)– 13(1,13)) line at 244.3 GHz, only in Fig. 12. A first-moment map of the SO 2(14(0,14)–13(1,13)) line is shown in Fig. 17 in the middle left panel. Although the source remains barely resolved (source size 0:4300) by the beam ( 0:69000:4900), there is a signature of a rotating structure in SO 2, as was also seen in EP Aqr (Homan et al. 2018b; Tuan-Anh et al. 2019). Inte- grated intensity maps of three SO 2lines (SO 2(9(3, 7)–9(2, 8)), which has the lowest upper level energy of the SO 2lines detected here ( Eu=64K); SO 2(14(0,14)–13(1,13)), the strongest line, and SO 2(34(4,30)–34(3,31)), which has the highest upper level energy of the detected lines, Eu=595K) are shown in Fig. B.4. All lines show kinematic structure in the E–W direction, approx- imately orthogonal to the outflow structure seen in CO and SiO, cf. Fig. 18. We derive the rotational temperature and column density of the SO 2-emitting region with a population diagram analysis (Sect. 3.6) and by an XCLASS modeling (Appendix D). Both methods give a similar rotational temperature of 320350K and a column density of 3:51015cm2. 3.2.7. PN We detect a line at 234.936 GHz that we ascribe to the PN molecule, which would be the first detection of PN in RS Cnc. PN has been detected in several M-type stars (e.g., De Beck et al. 2013; Ziurys et al. 2018), and in the C-rich envelopes of IRC +10216 and CRL 2688 (Guélin et al. 2000; Cernicharo et al. 2000; Milam et al. 2008). The presence of PN in an MS-type star therefore does not seem to come as a surprise. However, RS Cnc 5http://fruity.oa-teramo.inaf.it/ A135, page 8 of 27
10.1051_0004-6361_202141662
page_0007
J. M. Winters et al.: Molecules, shocks, and disk in the axi-symmetric wind of the MS-type AGB star RS Cancri Table 2. Parameters of the detected SO 2lines used for the population diagram analysis. Frequency WI=R S(v)dv g ulog10(Aul) Eu=kab (GHz) (Jy km s1) (s1) (K) (arcsec2) 214.6894 0.1410.0385 33 –4.0043 147.843 0.90 0.68 216.6433 0.166 0.0434 45 –4.0329 248.442 0.89 0.67 234.1871 0.1600.0439 57 –3.8401 403.033 0.71 0.56 244.2542 0.293 0.0698 29 –3.7855 93.901 0.69 0.49 245.5634 0.170 0.0451 21 –3.9240 72.713 0.69 0.49 248.0574 0.119 0.0333 31 –4.0939 119.328 0.69 0.48 258.3887 0.153 0.0396 65 –3.6773 531.100 0.63 0.45 258.9422 0.192 0.0524 19 –3.8800 63.472 0.64 0.45 259.5994 0.182 0.0448 61 –3.6835 471.496 0.63 0.45 263.5440 0.152 0.0448 61 –3.7227 459.038 0.61 0.42 265.4820 0.168 0.0448 69 –3.6426 594.661 0.61 0.42 Notes. Data are merged from A-configuration and D-configuration. Quoted errors include the rms errors of the Gaussian fits in the uvplane and the absolute flux calibration accuracy of 20%. The SO 2line parameters are retrieved from the CDMS and are based on the calculations by Lovas (1985) and Müller & Brünken (2005). Fig. 12. Profile of SO 2(14(0,14)–13(1,13)) with A-configuration and D- configuration merged, a spectral resolution of 3 km s1, and emission integrated over the central 200200aperture. Fig. 13. Profile of PN( N=5–4,J=6–5) with A-configuration and D- configuration merged, a spectral resolution of 3 km s1, and emission integrated over the central 200200aperture. appears to be the source with lowest mass-loss rate in which this molecule has been reported so far. The PN line profile is shown in Fig. 13. The line is spatially resolved at 0:800, which places it in about the same region as HCN and SO. The first-moment map of this line also shows signatures of rotation but due to the weak- ness of the line, the evidence is low. An integrated intensity map of PN is presented in Fig. B.5, showing that the line-emitting region is slightly spatially resolved. The 3feature seen about 1:500south of the phase center should not be considered as a detection but rather as a noise peak, as long as this structure is not confirmed by higher sensitivity observations. Fig. 14. Line wings in SiO(5–4) and SiO(6–5) compared to CO(2–1). The emission is integrated over the central 500500aperture. Fig. 15. High velocities close to the line of sight as seen in SiO. PV maps are shown in the Vzvs.Rplane for SiO(5–4) ( left) and SiO(6–5) (right ). The horizontal black line indicates the wind terminal velocity as traced in CO and the white scale bar indicates the spatial resolution. R=p (Dec)2+(RA)2,jVzj=jvlsrvlsr;j: 3.3. High-velocity wings in SiO, and in other molecules In SiO, five lines in three different vibrational states ( v=0,1,2) are detected (see Figs. 5 and 6). The vibrational ground-state lines clearly indicate the presence of material at velocities much higher than the wind terminal velocity of 8 km s1as traced by CO lines at this stellar latitude (see Sect. 4.1). This is illustrated in Fig. 14, and in Fig. 15 where we define vz=vlsrvlsr;, the Doppler velocity relative to the star. The high-velocity region is centered on the line of sight and is confined to the inner 0:300; see Fig. 15. A similar feature was seen in high-spatial- resolution observations of other oxygen-rich, low-mass-loss-rate A135, page 9 of 27
10.1051_0004-6361_202141662
page_0008
A&A 658, A135 (2022) AGB stars, such as W Hya (Vlemmings et al. 2017), EP Aqr (Tuan-Anh et al. 2019), oCet (Hoai et al. 2020), R Dor (Decin et al. 2018; Nhung et al. 2019a, 2021), and in 15 out of 17 sources observed in the ALMA Large Program ATOMIUM (Decin et al. 2020; Gottlieb et al. 2022), calling for a common mechanism causing high-velocity wings in this type of object. In the case of EP Aqr, where the bipolar outflow axis almost coincides with the line of sight (with an inclination angle of i10), the high- velocity wings were interpreted in terms of narrow polar jets. For R Dor and oCet, which do not show obvious signs of axial symmetry in their winds, such an interpretation could not be retained and it was argued instead that the high-velocity wings were caused by (a mixture of) turbulence, thermal broadening, and some effect of shocks, acting at distances below some 10 to 15 AU from the central star. The presence of broad wings in the SiO lines emitted from RS Cnc, whose symmetry axis is inclined by30with respect to the line of sight (see Sect. 4.1), lends sup- port to the latter type of interpretation and casts serious doubts on the polar jet interpretation proposed earlier for EP Aqr, which shows a morpho-kinematics similar to that of RS Cnc (Nhung et al. 2015b). Indeed, if the broad line widths are present regard- less of the orientation of a possible symmetry axis with the line of sight, they must be caused by a mechanism of nondirectional (accounting for the resolving beam) nature. A possible candi- date, whose action is limited to the close vicinity of the star, is pulsation-driven shocks that dissipate their energy relatively close to the star and imply positive and negative velocities in the shocked region that can be much higher than the terminal out- flow velocity of the wind. Such structures could be explained by the B-type models discussed in Winters et al. (2000b) as presented in Winters et al. (2002); see their Fig. 3. Recent 3D model calculations that self-consistently describe convection and fundamental-mode radial pulsations in the stellar mantle would provide the physical mechanism that leads to the development of such shocks close to the star surface (e.g., Freytag et al. 2017) and could therefore replace the simplified inner boundary condi- tion (the so-called “piston approximation”) that was used in the earlier 1D models mentioned above. In the data presented here, wings at high Doppler velocity are seen in nearly all lines detected with sufficient sensitivity to probe the profile over at least vlsr;10km s1. This is illus- trated in Fig. 16, where vzprofiles are integrated over a circle of radius 0:200centered on the star. Gaussian profiles centered at the origin are shown as visual references (not fits), showing how absorption produces asymmetric profiles. A major differ- ence is seen between vibrational ground-state lines, which have a Gaussian FWHM of 10km s1, and vibrationally excited- state lines, which have a Gaussian FWHM of 14km s1. Such a difference is not surprising, assuming that the high- velocity wings are formed in the inner layer of the circumstellar envelope (CSE), which is preferentially probed by the ( v=1) lines. In this context, we note that Rizzo et al. (2021) recently reported the detection of a narrow SiO( v=1, 1–0) maser line in RS Cnc at a velocity of +14 km s1with respect to the star’s lsr velocity. The effect of shocks on line profiles was first observed in the near-infrared range on CO ro-vibrational lines, probing the stellar photosphere and the innermost circumstellar region within10R(e.g.,Cyg, an S-type star, Hinkle et al. 1982). Very-high-angular-resolution observations obtained over the past decade using VLT, VLTI, and ALMA show that the effect of shocks from pulsations and convection cell ejections is confined within some 10 AU from the star (see, e.g., Khouri et al. 2018; Höfner & Olofsson 2018; Ohnaka et al. 2019, and references therein). Rotation, when observed, is instead found Fig. 16. Line profiles of different molecules on a logarithmic intensity scale. Gaussian profiles are shown for comparison, FWHM =10km s1 for the ground-state lines of all molecules, and FWHM =14km s1 for the (v=1) lines of SiO. All observed profiles are integrated over R<0:200. to extend beyond this distance, typically up to 20 AU (e.g., Vlemmings et al. 2018; Homan et al. 2018a; Nhung et al. 2021). The angular resolution of the present data is insufficient to detect such differences directly; however, the effect of rotation and shocks on lines of sight contained within a beam centered on the star depends on the region probed by each specific line: lines that probe the inner layers exclusively, such as the ( v=1) lines, are mostly affected by shocks, and somewhat by rotation; CO lines, for which the probed region extends very far out, see little effects of rotation and even less effects of shocks because the emission from the inner envelope provides too small a fraction of the total emission. Between these two extremes, the relative importance of the contributions of shocks and rotation depends on the radial extent of the region probed by the line. Such an interpretation is consistent with the data displayed in Fig. 16. 3.4. Rotation In Fig. 17, we present first-moment maps of HCN(3–2) (left), SO2(14(0,14)–13(1,13)) (middle left), SiO( v=1, 6–5) (middle right), and SO(7(6)–6(5)) (right). At projected distances from the star not exceeding 0:500, all four tracers display approximate anti-symmetry with respect to a line at PA10. This is sugges- tive of the presence of rotation in the inner CSE layer around an axis that projects on this line in the plane of the sky. Such a morpho-kinematic structure has also been observed in other stars, notably R Dor (Vlemmings et al. 2018; Homan et al. 2018a; Nhung et al. 2021). The angular resolution of the present data does not allow for a detailed exploration of this region, which prevents us from commenting on its possible cause. Neverthe- less, the anti-symmetry axis of the velocity pattern projected on the plane of the sky at a PA that approximately coincides with the projected symmetry axis of the polar outflows (see Sect. 4.1) A135, page 10 of 27
10.1051_0004-6361_202141662
page_0009
J. M. Winters et al.: Molecules, shocks, and disk in the axi-symmetric wind of the MS-type AGB star RS Cancri Fig. 17. First-moment maps of different lines, indicating a possibly rotating structure (see Sect. 3.4). Left: HCN(3–2), middle left : SO 2(14(0,14)– 13(1,13)), middle right : SiO(v=1,6–5), right : SO(7(6)–6(5)). The black ellipses indicate the synthesized beam. is remarkable and suggests that rotation is taking place about this same polar axis in the inner CSE layer. The line-of-sight velocities of these structures are small, on the order of the velocities derived from CO for the equatorial region, and we interpret them here as possible signs of rotation (rather than indicating another bipolar outflow oriented perpen- dicular to the larger scale outflow traced in CO and SiO ( v=0) lines). We note that out of these four lines, the HCN(3–2) line is detected with the highest S/N (S/N =233in the line peak, cf. Table 1). The mean Doppler velocity hvzi, averaged over the inner 0:500, of the HCN line can be fit in position angle !, measured counter-clockwise from north, by hvziHCN=0:19 km s1+1:0 km s1sin(!19); (1) whereas the SiO( v=1, 6–5) velocity is well fit by hvziSiO (v=1;65)=0:37 km s1+0:46 km s1sin(!26):(2) The small offsets of 0:3km s1on average are within the uncertainty attached to the measurement of the star’s LSR velocity. The coefficients of the sine terms measure the pro- jected rotation velocity, namely the rotation velocity divided by the sine of the angle made by the rotation axis with the line of sight. Assuming that the rotation axis is the axi-symmetry axis of the CSE, this angle is i30(see Sect. 4.1), meaning rotation velocities of2and1km s1for HCN and SiO respectively. Observations of higher angular resolution are needed to confirm the presence of rotation within a projected distance of 0:500from the star and we prefer to summarize the results presented in this section in the form of an upper limit to the mean rotation velocity of a few km s1. 3.5. Global outflow structure traced by CO and SiO The detailed structure of the morpho-kinematics of the CSE has been studied using observations of the12CO(1–0) and12CO(2–1) molecular line emission. The analyses of Hoai et al. (2014) and Nhung et al. (2015b) confirmed the interpretation of the two- component nature of the Doppler velocity spectrum originally given by Libert et al. (2010). The CSE is axi-symmetric about an axis making an angle of i30with the line of sight and projecting on the plane of the sky at a position angle !7east of north (see also the sketch in Fig. 4). The expansion velocity reaches8to9km s1along the axis – we refer to this part of the CSE as bipolar outflow – and 3to4km s1in the plane perpendicular to the axis – we refer to this part of the CSE as equatorial enhancement. The transition from the equator to thepoles of the CSE is smooth. Section 4.1 below, using observa- tions of the12CO(2–1) and13CO(2–1) molecular lines, confirms and significantly refines this picture. The right panels of Fig. 20 show projections of the CSE on the plane containing the axis and perpendicular to the plane of the sky, which give a good qualitative idea of the global structure. Velocity-integrated channel maps of the CO(2–1) and SiO(6–5) observations analyzed in the present article are dis- played in Fig. 18. They clearly show the bipolar outflows, inclined toward the observer in the north and receding in the south. We note that the red wings are brighter than the blue wings as a result of absorption (see Sects. 4.1 and 4.2) The SiO- emitting region is seen to be significantly more compact than the CO-emitting region; this is in conformity with observations of many other oxygen-rich AGB stars and is generally interpreted as the result of SiO molecules condensing on dust grains and being ultimately dissociated by the interstellar radiation at some 200 AU from the star, well before CO molecules are dissociated (see e.g., Schöier et al. 2004). 3.6. Temperature and SO 2abundance In this section, we use the 11 detected SO 2lines to derive an approximate temperature and column density of the SO 2- emitting region by means of a population diagram. Following Goldsmith & Langer (1999), in the optically thin case, the col- umn density of the upper level population Nuof a transition u->l can be expressed as Nu=8k2 hc3AulZ Tbdv: (3) Nuis the column density of the upper level population of the transition, kandhare the Boltzmann and Planck constant, respectively, is the line frequency, cthe speed of light, Aulis the Einstein coefficient for spontaneous emission of the transi- tion, andR Tbdvis the velocity-integrated main-beam brightness temperature. The latter is converted to the surface brightness distribution of the source Sper beam, measured by the inter- ferometer, by means of Tb=2 2k bS; (4) where=c is the observing wavelength, and b=ab 4 ln 2witha andbbeing the major and minor axis of the synthesized beam. We determineR S(v)dv=:WIfrom a circular Gaussian fit to the velocity-integrated emission in the uv-plane, where the inte- gration is taken from (vlsr,*4:5)km s1to(vlsr;+4:5)km s1, that is over the three central channels of the SO 2lines. A135, page 11 of 27
10.1051_0004-6361_202141662
page_0010
A&A 587, A91 (2016) about their formation paths during the past evolution of the core. In particular, their degree of deuterium fractionation is a relic of the conditions that were prevailing in earlier, colder stages (e.g.Taquet et al. 2012 ,2014 ;Aikawa et al. 2012 ). Apart from methanol, relativ ely little is known about deuter- ation of complex organic molecules in the interstellar medium,in particular in high mass star-forming regions. A few detec- tions or tentative detections of deuterated (complex) organic molecules have been reported toward Orion KL ( Gerin et al. 1992 ;Daly et al. 2013 ;Esplugues et al. 2013 ;Neill et al. 2013 ; Coudert et al. 2013 ), but no systematic study of the deuteration of complex organic molecules in high mass star-forming regionshas been reported so far. Sagittarius B2, hereafter Sgr B2, is one of the most mas- sive star-forming regions in the Galaxy. It is located in the cen-tral molecular zone, close to the Galactic center. Its current star formation rate qualifies Sgr B2 as a mini-starburst region (see, e.g., Appendix A of Belloche et al. 2013 ). The cloud contains several sites of ongoing, high-mass star formation. One of these sites, Sgr B2(N), contains two hot molecular cores that we des- ignate as Sgr B2(N1) and Sgr B2(N2) 2. Their angular separation is 5/prime/primein the north-south direction, corresponding to 0.2 pc in projection at a distance of 8.3 kpc ( Reid et al. 2014 ). Their ve- locities projected along the line of sight di ffer by 9–10 km s−1. Both are characterized by extremely high H 2column densities (>1025cm−2over few arcsec, see Belloche et al. 2008 ,2014 ; Qin et al. 2011 ). They are both in an early stage of star forma- tion when a (massive) protostar has already formed and startedto heat up its circumstellar envelope. The high kinetic temper- atures of the hot cores ( ∼150–200 K) lead to the sublimation of molecules that formed in the ice mantles of dust grains dur-ing the prestellar phase and the warming-up period of the proto- stellar phase. As a result of both the high temperatures and col- umn densities, numerous complex organic molecules have beendetected toward Sgr B2(N), many of these for the first time in the interstellar medium, since the beginning of radio astronomy nearly five decades ago. Following up a molecular line survey of Sgr B2(N) per- formed with the IRAM 30 m telescope that led to the first detection in space of a few new complex organic molecules(Belloche et al. 2008 ,2009 ,2013 ), we performed a system- atic line survey of Sgr B2(N) in the 3-mm atmospheric win- dow at high angular resolution and sensitivity with the Atacama Large Millimeter/submillimeter Array (ALMA) in its Cycles 0 and 1. This survey is called EMoCA, which stands for Exploring Molecular Complexity with ALMA, and it aims to investigatemolecular complexity in the interstellar medium. One of the ini- tial results of EMoCA was the first interstellar detection of a branched alkyl molecule ( Belloche et al. 2014 ). Here, we take advantage of this sensitive survey to explore, for the first time in a systematic way, the deuterium fractionation of complex or- ganic molecules in Sgr B2(N2). We focus on Sgr B2(N2) ratherthan the main hot core Sgr B2( N1) because the former has rel- atively narrow linewidths ( ∼5k ms −1) at the angular resolution of EMoCA (∼1.8/prime/prime), while the latter still has prominent linew- ings like in our previous single-dish survey. A companion paper reports on the detection of alkanols and alkanethiols based on EMoCA ( Müller et al. 2016b ). The article is structured as follows. The observational setup and the process of data reduction are described in Sect. 2. Section 3explains the method employed to model the observed 2They were named P1 and P2 in Belloche et al. (2008 )a n dS M A 1a n d SMA2 in Qin et al. (2011 ).spectra in the approximation of l ocal thermodynamic equilib- rium (LTE) and Sect. 4gives some details about the spectro- scopic predictions used to generate the synthetic spectra. Theresults of the analysis are reported in Sect. 5and discussed in Sect. 6. Section 7gives our conclusions about deuterium fractionation of complex organic molecules in Sgr B2(N2). 2. Observations and data reduction 2.1. Observations We used ALMA to perform a complete spectral line survey to-ward Sgr B2(N) between 84.1 and 114.4 GHz. The field was cen- tered at (α,δ) J2000=(17h47m19.87s,−28◦22/prime16/prime/prime), halfway be- tween Sgr B2(N1) and (N2), which are separated by 4 .9/prime/primein the north-south direction. The size (HPBW) of the primary beam of the 12 m antennas varies between 69/prime/primeat 84 GHz and 51/prime/prime at 114 GHz ( Remijan et al. 2015 ). The spectral line survey was divided into five spectral setups. Each setup was observed in one polarization and delivered fourspectral windows, two per sideba nd. The separation between the centers of the lower and upper sidebands is 12 GHz. Each spec- tral window has a bandwidth of 1875 MHz and a channel spacingof 244.141 kHz, but the spectra were smoothed to a spectral res- olution of 488.3 kHz (1.7 to 1.3 km s −1). Each pair of adjacent spectral windows has an overlap of about 50 MHz. Details aboutthe frequency coverage, the date of observation, the number of antennas, the range of baselines, the on-source integration time, and the bandpass, amplitude, and phase calibrators are given inTable 1. Setups S1 to S4 were observed in Cycle 0 while setup S5 was observed in Cycle 1. As reported in Table 1, setups S1 and S5 were observed only once, but setups S2, S3, and S4 wereobserved on several days, between two and four times each. 2.2. Data reduction The data was calibrated and imaged with the CommonAstronomy Software Applications package (CASA). We used version 4.2.0 (r28322) for setups S1 to S4 and version 4.2.1 (r29047) for setup S5. We used the standard procedures providedby the Joint ALMA Observatory to apply the bandpass, ampli- tude, and phase calibrations. The deconvolution was performed with the csclean imager mode and a Briggs weighting scheme with a robust parameter of 0.5. The cell size was set to 0 .3 /prime/prime. In addition, three or four iterations of self-calibration were per- formed using a strong spectral line detected toward Sgr B2(N1) in each setup. This significantly improved the dynamical range in the resulting images. The spectra toward Sgr B2(N1) and (N2) are full of lines and close to the confusion limit. It is thus di fficult to separate the line emission from the continuum emission in a systematic way forthe full data cubes, but it is a necessary step to produce sepa- rate line and continuum maps. For each spectral window of each setup, we selected six groups of a few channels that seemed to befree of strong line emission. A first-order baseline was fitted to these selected channels and the result of the fit was used to split each data cube into two cubes: one for the line emission and onefor the continuum emission. Given the di fference in systemic ve- locity between the two hot cores ( ∼10 km s −1,s e e Belloche et al. 2013 ), we selected different sets of channels for the northern and southern parts of the field. This process of baseline subtraction was performed with the CLASS software3. 3Seehttp://www.iram.fr/IRAMFR/GILDAS A91, page 2 of 66
10.1051_0004-6361_201527268
page_0001
A. Belloche et al.: Deuterated complex organic molecules in Sgr B2(N2) Table 1. Observational setups of the EMoCA survey. Setup Frequency range Date of tstartaNabBaseline tintdCalibratorsePeakf LSB USB observation (UTC) rangecBA P ΔαΔδ GHz GHz yyyy-mm-dd hh:mm m min/prime/prime /prime/prime S1 84.1–87.8 96.2–99.8 2012-08-27 01:05 26 17–400 54.7 1 3 2 0.2 −2.4 S2 87.7–91.4 99.7–103.4 2012-06-05 09:03 20 14–392 24.2 1 3 2 1.2⋆−2.4 2012-07-01 07:04 13 25–393 40.0 1 3 2 1.2⋆−2.6 2012-07-03 06:36 21 14–395 48.4 1 3 2 0.6⋆−2.3 2012-09-28 21:57 25 20–387 44.1 1 4 2 0.2 −2.5 S3 91.4–95.1 103.4–107.1 2012-06-06 08:20 18 15–395 40.2 1 3 2 0.7⋆−2.4 2012-06-18 07:29 22 15–395 40.4 1 3 2 0.6⋆−2.2 S4 95.0–98.7 107.0–110.7 2012-07-04 05:38 21 17–398 8.1 1 3 2 0.1 −2.3 2012-08-01 02:32 24 19–442 34.9 1 3 2 0.1 −2.4 2012-08-10 00:45 26 21–400 35.0 1 3 2 0.2 −2.4 S5 98.7–102.4 110.7–114.4 2014-04-05 06:22 38 15–413 24.4 2 4 5 0.2 −2.4 Notes.(a)Start time of observation.(b)Number of ALMA 12 m antennas.(c)Minimum and maximum projected baseline separations.(d)On-source integration time.(e)Bandpass (B), amplitude (A), and phase (P ) calibrators. The calibrators are: 1: B1730-130, 2: J 1700-2610, 3: Neptune, 4: Titan, 5: J1744-3116.(f)Offset of the continuum peak position of Sgr B2(N1) with respect to the phase center, in equatori al coordinate system (J2000). Measurement sets with Δαdiffering from 0.1/prime/primeby more than 0.2/prime/prime(marked with a⋆) are believed to be a ffected by an astrometric problem. We checked the accuracy of the relative astrometry between the ten measurement sets by fitting the peak position of thecontinuum emission toward Sgr B2(N1) in selected channels of the line+continuum data cubes that appeared to be free of line emission. It turns out that the first three measurement sets ofsetup S2 and both measurement sets of setup S3 are a ffected by an astrometric problem: the continuum peak of Sgr B2(N1) is shifted by 0.6 /prime/primeto 1.2/prime/primein right ascension with respect to all other measurement sets (see Table 1). The dispersion of the peak position in declination is also a bit higher for the af- fected measurement sets compared to the nona ffected measure- ment sets. The average peak position of Sgr B2(N1) in all non- affected measurement sets is at ( Δα,Δδ)=(0.15/prime/prime,−2.40/prime/prime), i.e. (α,δ)J2000=(17h47m19.881s,−28◦22/prime18.40/prime/prime). The five affected measurement sets were obtained after transit when the sourcewas setting and the phase calibrator was at low elevation, which leads us to believe that the shift of the a ffected measurement sets may be due to an inacurrate calibration of the atmospheric phasefluctuations. As a result, we ignored the a ffected measurement sets of setup S2 and used only its fourth measurement set. For setup S3, both measurement sets were used but the o ffset was approximately compensated for by modifying the visibilities of the phase calibrator with the CASA task fixvis before the phase calibration. After this correcti on, the relative positional accuracy of all measurement sets selected for this work is on the order of±0.1 /prime/primein both right ascension and declination. The measurement sets of setup S3 were merged into one sin- gle measurement set with the CASA task concat before imag- ing. The same was carried out for setup S4. Only one measure- ment set was used for the other three setups. The size (HPBW)of the synthesized beam and the rms noise level in the final cubes are given in Table 2. The noise level of each spectral win- dow corresponds to the median of the noise levels measured inall channel maps using the procedure go noise in GREG 3.The noise levels reported in Table 2were measured in the continuum- subtracted datacubes. They are typically a factor ∼2 higher in the line+continuum datacubes. Based on the redundancies of the measurement sets and the spectral overlap between the setups, we estimate the relative cal-ibration uncertainty on the flux density to be on the order of 15%.3. Radiative transfer modeling of the line survey We used the input parameters of our LTE model of theIRAM 30 m spectrum of Sgr B2(N) ( Belloche et al. 2013 )a s a starting point to assign the lines detected in the ALMA spec- tra and model the emission of the detected molecules. Given the high H 2densities of Sgr B2(N1) and (N2) ( ∼108cm−3at arc- second scale, Belloche et al. 2008 ,2014 ;Qin et al. 2011 ), the LTE approximation is appropriate. We used the software Weeds (Maret et al. 2011 ) to produce synthetic LTE spectra that take into account the radiative transfer and the (spectral-window- and measurement-set-dependent) angular resolution of the observa- tions. We performed the modeling for each species separately,and then we linearly added the contributions of all detected species to the emitted spectra to obtain the final synthetic spec- trum (hereafter called the full LTE model). This approximation is valid for optically thin lines that overlap in frequency space orfor (optically thick or thin) lines of species that are emitted from separated regions within the beam, but it is no longer correct for frequency-overlapping optically thick lines of species that arecospatial or aligned along the line of sight. In such cases, the synthetic spectrum overestimates the actual line flux density. The model of each species is defined by five parameters: angular size of the emitting region assumed to be Gaussian, column density, rotational temperature, velocity o ffset with re- spect to the assumed systemic velocity of the source, and linewidth (FWHM). For a given species, the source size was de- rived from two-dimensional Gau ssian fits to the integrated in- tensity maps of all transitions that were well detected and found to be free of contamination (based on the full LTE model). The source size was set to the median deconvolved size of all suchtransitions. The other four parameters were optimized by eye. We constructed population diagrams based on the transitions that are well detected and not severe ly contaminated by transitions of other species. In the case where a transition was partially con- taminated, the contributions of the contaminating species was re- moved from the measured integrated intensities, on the basis ofthe full LTE model. Each population diagram was also corrected for optical depth following the method described in Goldsmith & Langer (1999 ), using the opacities delivered by Weeds. We used the population diagrams to verify that the rotational temperature A91, page 3 of 66
10.1051_0004-6361_201527268
page_0002
A&A 587, A91 (2016) Table 2. Beam sizes and noise levels. Setup SPWaFrequency Synthesized beam rmsc range HPBW PAbmJy MHz/prime/prime×/prime/prime ◦beam−1K S1 0 84 091–85 966 2 .1×1.5−85 3.0 0.16 1 85 904–87 779 2 .0×1.5−83 2.7 0.14 2 96 154–98 029 1 .8×1.4−85 3.0 0.16 3 97 904–99 779 1 .8×1.3−85 3.1 0.16 S2 0 87 729–89 604 1 .9×1.7 86 3.1 0.15 1 89 554–91 429 1 .8×1.6 52 2.8 0.15 2 99 728–101 602 1 .6×1.4 48 2.7 0.14 3 101 552–103 427 1 .6×1.4 49 2.7 0.14 S3 0 91 368–93 242 2 .9×1.5 84 3.4 0.12 1 93 193–95 067 2 .8×1.5 83 3.1 0.10 2 103 365–105 239 2 .5×1.3 82 3.4 0.11 3 105 189–107 064 2 .5×1.3 82 3.6 0.12 S4 0 95 021–S96 896 1 .9×1.4−82 1.9 0.10 1 96 846–98 720 1 .8×1.3−82 1.9 0.10 2 107 019–108 893 1 .7×1.2−83 2.2 0.11 3 108 843–110 718 1 .6×1.2−82 2.3 0.12 S5 0 98 672–100 546 1 .8×1.4−76 2.8 0.14 1 100 496–102 370 1 .7×1.4−76 2.7 0.13 2 110 669–112 543 1 .6×1.3−72 3.5 0.17 3 112 494–114 368 1 .6×1.2−77 4.9 0.24 Notes.(a)Spectral window.(b)Position angle east from north.(c)Median rms noise level measured in the channel maps of the continuum-removed data cubes. derived in the course of the (manual) modeling with Weeds made sense. In the population diagrams corrected for optical depth and contamination, the residual dispersion of the synthetic data-points (red crosses) results in part from the frequency boundaries set to integrate the intensity. These boundaries are a compromise between covering the line as much as possible and limiting thecontamination from other species emitting at nearby frequencies as much as possible. In addition, the correction for optical depth is an approximation and may also introduce some bias. Finally,another limitation of this fit is that it can be biased by residualcontamination that remains even after removal of the contribu- tion of the identified contamina ting species. Therefore, we be- lieve that the formal errors on the rotational temperature derivedfrom the fit to the population diagrams do not necessarily rep- resent the true errors on this temperature and should be viewed with caution. The emission of vibrationally excited states of a given molecule were modeled independentely of the vibrational ground state. The emission of isotopologues of a given moleculewere also modeled separately. The physical structure of the source assumed for the model- ing is uniform. This may sound simplistic given that temperature and density gradients are expect ed in the envelope of Sgr B2(N2) (e.g., Rolffs et al. 2011 ). It turns out that, even with such a sim- ple assumption, the spectra of most complex organic molecules detected toward Sgr B2(N2) can be well reproduced so we areconfident that the rotational temperatures and column densities derived from our analysis are reliable. In the following, we count a line of a given species as a de- tected line if its frequency, peak intensity, and width are wellreproduced by our model and the line is not (or barely) con- taminated by emission from other species. As a counter exam- ple, a synthetic line that is consistent with the observed spec-trum, i.e., that has a peak intensity simply below the intensity ofthe detected signal, but would still remain consistent if it were shifted by a frequency o ffset typically equal to its linewidth is not counted as detected. We emphasize that our definition of a de- tected line is very conservative but we believe that it is requiredto avoid unsecure molecule detections. The complete list of transitions identified in our survey is presently not available but the list of lines identified in our pre- vious single-dish survey of Sgr B2(N) can be taken as reference (Belloche et al. 2013 ). 4. Spectroscopic predictions The origin of the spectroscopic predictions used to model the emission of the species reported in Sect. 5is provided here. Predictions for the three singly deuterated species of ethyl cyanide were taken from the catalog of the Cologne Databasefor Molecular Spectroscopy (CDMS 4,Müller et al. 2001 ,2005 ; tags 56 509, 56 510, and 56 511, all version 1). They are based onMargulès et al. (2009 ). All other ethyl cyanide data were also taken from the CDMS. The main species predictions are based onBrauer et al. (2009 ) with published data in the range of our survey from Fukuyama et al. (1996 ). Transition frequencies of the isotopologues containing one13C were taken from Richard et al. (2012 ), those for the15N isotopologue from Margulès et al. (2009 ). Vibrational corrections to the rotational partition func- tion, and thus to the column density, were derived for the main isotopologue from Heise et al. (1981 ) and applied to all species. Only limited isotopic data are available. It is safe to assume thatdifferences among the isotopologues are small, most likely not exceeding a few percent because of the large number of heavy atoms in the molecule. Predictions for singly deuterated methyl cyanide were taken from the CDMS catalog (tag 42 511, version 2). This entry is 4Seehttp://www.cdms.de A91, page 4 of 66
10.1051_0004-6361_201527268
page_0003
A. Belloche et al.: Deuterated complex organic molecules in Sgr B2(N2) based on Nguyen et al. (2013 ). Predictions for CH 3CN in its /v18=1 and 2 excited states are based on preliminary results from Müller et al. (2015 ), those in /v14=1 are preliminary data from a subsequent study (Müller et al., in prep.). Transition frequen- cies in the range of our study are from Bauer & Maes (1969 ) andBauer (1971 )f o r /v18=1 and 2, respectively. Predictions for isotopic species with15N or one or two13C in their ground vibra- tional states are from the CDMS. They are based on Müller et al. (2009 ) with transition frequencies in the range of our survey from Demaison et al. (1979 ). Predictions for13C isotopologues in their /v18=1 states are based on preliminary data from Müller et al. (2016a ). Vibrational corrections to the partition function were included in the private entries for the main isotopic species.T h e ya r en o wa v a i l a b l ei n Müller et al. (2015 ). As the correc- tions are small, the error using the values from the main isotopic species for the other isotopologues as well is very small, evenfor CH 2DCN. Predictions for deuterated vinyl cyanide CH 2CDCN were taken from the molecular spectroscopic database of the Jet Propulsion Laboratory (JPL5,Pickett et al. 1998 ; tag 54 004, ver- sion 2). We prepared predictions for cis-CHDCHCN and trans - CHDCHCN. All predictions are based on Colmont et al. (1997 ). Predictions for C 2H3CN and isotopic species with one13Co r with15N were taken from the CDMS and are based on Müller et al. (2008 ). Transition frequencies in the range of our survey are mostly from that study. For the main species, they are, toa large extent, also from Baskakov et al. (1996 ). Predictions for excited states of vinyl cyanide used in the present work are based on Cazzoli & Kisiel (1988 ) and unpublished data from one of us (HSPM). These data included vibrational corrections which are essentially complete at 200 K. The vibrational ener- gies were gathered from several sources. A recent compilation of low-lying vibrational states is available in Kisiel et al. (2015 ). It is safe to assume that di fferences among the isotopologues are small, most likely not exceedin g a few percent because of the large number of heavy atoms in the molecule. Predictions for deuterated c yanoacetylene were taken from the CDMS catalog (tag 52 508, version 2). This entry is based on Spahn et al. (2008 ). All other cyanoacetyl ene predictions were also taken from the CDMS. The /v1 7=1 predictions of the main species are based on Thorwirth et al. (2000 ) with data in the range of our survey from Yamada & Creswell (1986 ). All pre- dictions of isotopologues containing one or two13Ca r eb a s e d onThorwirth et al. (2001 ), and those for HC15 3No n Fayt et al. (2004 ). Ground state transition frequencies for singly substituted species in the range of our survey were taken from Creswell et al. (1977 ). Vibrational contributions to the partition functions of HC 3Na n dD C 3N can be evaluated from the study of their low-lying vibrational states by Uyemura et al. (1982 ). Isotopic shifts, in particular of the lowest ν7mode, are much smaller for 13Co r15N species. Therefore, using vibrational corrections of the main isotoplogue introduces small errors for these species. Predictions for deuterated methanol CH 2DOH were taken from the JPL catalog (tag 33 004, version 1). They are based on Pearson et al. (2012 ) with rest frequencies almost entirely from that study. With the use of torsional energies from Lauvergnat et al. (2009 ), we estimate a vibrational correction factor to the partition function of 1.15 at 160 K. For CH 3OD, we prepared a catalog entry based on Anderson et al. (1988 ), with frequen- cies updated to the values published in Duan et al. (2003 ). We estimated the partition function to be 11 770 at 150 Kand 25 550 at 225 K, taking torsional energies of CH 3OD in 5Seehttp://spec.jpl.nasa.govAnderson et al. (1988 ) into account. Details on other methanol isotopologues are given in Müller et al. (2016b ). Predictions for all singly deuterated species of ethanol were taken from the CDMS (tags 47 515 to 47 518, all version 1). They are based on Walters et al. (2015 ) with rest frequencies almost entirely from that study. All other ethanol analyses weretaken from Müller et al. (2016b ), and further details can be found there. Conformational and vibrational corrections to the parti- tion function were taken from the main isotopologue for whichonly data were available. This assumption is reasonable, though errors may not be completely negligible. They are, however, difficult to evaluate. Predictions for both conformers of CH 2DOCHO were ex- tracted from Table 7 of Coudert et al. (2013 ) and split into two separate entries. The partition function was taken from theirTable 6. It is identical for both entries. Contrary to the CDMS en- tries for CH 2DCH 2CN and CH 2DCH 2OH, this treatment means that the two entries represent a single species with a statisticaldistribution (the out-of-plane conformer being twice as abun- dant as the in-plane one). This means that the column density derived for each conformer represents the total column densityof the molecule. However, we assumed a statistical distribu- tion (2:1) to compute and report individual column densities in Sect. 5. Predictions for the main isotopologue were taken from the JPL catalog. The entry is based on Ilyushin et al. (2009 ). Vibrational corrections to the partition function were derived from the 13C species ( Favre et al. 2014 ). These authors pro- vide rotational partition function values at di fferent temperatures as well as detailed vibrational corrections that are complete up to 300 K. The correction factors are 1.59 and 1.23 at 150 K forthe deuterated and main isotopic species, respectively. These val- ues differ because values for the deuterated species refer to the ground state only whereas contributions of the first excited stateswere already included for the main isotopologue. 5. Results In this section, we report the detection or tentative detectionof deuterated complex organic molecules toward Sgr B2(N2).Column density upper limits are also reported for several non- detections. Each subsection first presents the LTE model de- rived for the main isotopologue and its 13Ca n d/or15N isotopo- logues. This model is then used to obtain constraints on the col- umn density of the deuterated species. The rotational tempera- tures derived from fits to the population diagrams are reportedin Table 3and the parameters of the LTE model used to fit the spectra are listed in Table 4. The analysis toward Sgr B2(N2) was performed at the o ffset position (Δα,Δδ)=(−0.1 /prime/prime,2.6/prime/prime), i.e. (α,δ)J2000=(17h47m19.86s,−28◦22/prime13.4/prime/prime). 5.1. Deuterated ethyl cyanide CH 2DCH 2CN and CH3CHDCN About 154, 54, 38, and 37 lines of ethyl cyanide and its singly substituted13C isotopologues,13CH 3CH 2CN, CH 313CH 2CN, and CH 3CH 213CN, respectively, are detected toward Sgr B2(N2) (Figs. A.1–A.4). The15N isotopologue is also detected un- ambiguously, with nine detected lines (Fig. A.5). The best-fit LTE model fits very well all detected transitions, except the very optically thick lines of the main isotopologue, which itsignificantly underestimates. We ignored the lines with opti- cal depth higher than ∼2.5 to construct the population diagram of this species (Fig. A.6), while all lines of the 13C isotopo- logues that matched the criteria defined in Sect. 3were used A91, page 5 of 66
10.1051_0004-6361_201527268
page_0004
A&A 587, A91 (2016) Table 3. Rotational temperatures derived from population diagrams of selected (complex) organic molecules toward Sgr B2(N2). Molecule StatesaTfitb (K) CH 3CN /v18=1,/v18=2,/v14=1 253 (15) 13CH 3CN /v1=0,/v18=1 168 (13) CH 313CN /v1=0,/v18=1 165.5 (3.3) CH 2DCN /v1=0 136 (14) C2H5CN /v1=0 137.3 (1.6) 13CH 3CH 2CN /v1=0 138.3 (7.5) CH 313CH 2CN /v1=0 112 (11) CH 3CH 213CN /v1=0 150 (40) C2H3CN /v1=0,/v111=1,/v115=1,/v111=2 199.5 (3.4) 13CH 2CHCN /v1=0 255 (101) CH 213CHCN /v1=0 140 (31) CH 2CH13CN /v1=0 278 (126) H13CCCN /v1=0,/v17=1 171.1 (3.2) HC13CCN /v1=0,/v17=1 167.7 (5.5) HCC13CN /v1=0,/v17=1 177 (18) CH 3OCHO /v1t=0,/v1t=1 142.4 (4.4) Notes.(a)Vibrational states that were taken into account to fit the popu- lation diagram.(b)The standard deviation of the fit is given in parenthe- ses. As explained in Sect. 3, these uncertainties should be viewed with caution. They may be underestimated. for their population diagrams (Figs. A.7–A.9). The results of the linear fit to the population diagrams of all four isotopologues are given in Table 3. The rotational temperature is well con- strained to∼140 K for both C 2H5CN and13CH 3CH 2CN. This value is consistent with the result of the fit for the two other iso- topologues within the uncertainties. The temperature derived in this way depends on the model used to make the opacity correc-tion. With an earlier best-fit model that assumed a temperature of 170 K (instead of 150 K here), the fit to the population di- agrams of both C 2H5CN and13CH 3CH 2CN yielded a tempera- ture of∼150 K. This is the reason why we decided to assume a rotational temperature of 150 K for ethyl cyanide and all its isotopologues. The median source size derived for the selected lines of C2H5CN is about 1.15/prime/prime, but there seems to be a trend of de- creasing size with increasing upper level energy, from ∼1.3/prime/primeat low energy to∼0.8/prime/primeatEu∼700 K (Fig. 1). Similar results are obtained for13CH 3CH 2CN, with a median size of ∼1.25/prime/primeand a hint of a decrease down to ∼1.0/prime/primeatEu∼120 K. Our model does not treat such gradients. As a compromise, we used a source size of 1.2/prime/prime. With this source size and a rotational temperature of 150 K, we obtain an excellent fit to all emission lines of the four iso- topologues, except for the very optically thick lines of C 2H5CN (τmax∼60), as mentioned above. A better fit to these lines would be obtained by increasing the temperature and /or assuming a larger source size. Increasing the size to 1 .4/prime/primeturns out to be insufficient. A larger size would be inconsistent with the mea- sured sizes. Increasing the temperature to 200 K and the size to 1.3/prime/primeyields peak temperatures of the optically thick lines sim- ilar to those observed, but the synthetic lines look too saturatedcompared to the observed lines, and the fit to the optically thin lines becomes worse; lines with high upper level energies be- come overpredicted. A more complex model with nonuniformphysical parameters would probably be needed to reproduce the intensity and shape of the very optically thick lines. Assuming the same source size, rotational temperature, linewidth, and velocity o ffset as derived for C 2H5CN and its13Ca n d15N isotopologues, we looked for emission of the singly deuterated isotopologues, CH 3CHDCN and CH 2DCH 2CN. The former is a chiral molecule b ecause the carbon atom in the middle of the chain is linked to four di fferent atoms or func- tional groups. Both isotopologues are tentatively detected to- ward Sgr B2(N2) with 1 and 2 line(s), respectively (Figs. A.10 andA.11 ), the latter isotopologue in its out-of-plane conforma- tion only. For the in-plane confomer of CH 2DCH 2CN, we derive an upper limit only. This upper lim it is uncertain because the ap- parent inconsistency between the synthetic spectrum and the ob-served one around ∼101 190 MHz may result from a slight over- estimate of the baseline, at the 3 σlevel (Fig. A.12 ). Owing to the limited signal-to-noise ratios, the source size derived from theintegrated intensity maps of the uncontaminated lines assigned to CH 3CHDCN and CH 2DCH 2CN is uncertain, varying between unresolved and∼2/prime/prime. The emission looks compact in the maps. Thus, assuming the same source size as the other isotopologues looks reasonable. 5.2. Deuterated methyl cyanide CH 2DCN Methyl cyanide is clearly detected in its vibrational ground statetoward Sgr B2(N2) but its transitions are very optically thick (τ max∼50) and cannot be properly fitted in the framework of our simple model (Fig. A.13 ). Transitions from within its vibra- tionally excited states /v18=1a n d /v18=2 are also clearly detected, withτmax∼2.7 and 0.3, respectively (Figs. A.14 andA.15 ). We also find four transitions from within /v14=1 around 91 520 MHz and 109 820 MHz ( τmax∼0.06), but they partially su ffer from blends with other species (Fig. A.16 ). The assignment looks reasonable, but the detection should be considered tentative. The singly substituted13C isotopologues are very well de- tected, both in their vibrational ground state and in their first vibrationally excited state /v18=1( F i g s . A.17 –A.20 ). The fit to their population diagrams yields rotational temperatures of about 170 K (see Table 3and Figs. A.21 andA.22 ). The analysis of the integrated intensity maps of the13C isotopologues deliv- ers a source size of ∼1.4/prime/prime. For the main isotopologue, it seems that the source size decreases with the vibrational energy ( ∼1.2/prime/prime for/v18=1a n d∼0.8/prime/primefor/v18=2). With the assumption of a source size of 1 .4/prime/primeand a temper- ature of 170 K, our LTE modeling yields excellent and consis- tent fits to the13C isotopologues (both /v1=0a n d /v18=1) and to the /v18=1 transitions of the main isotopologue. However, it was necessary to increase the column density and linewidth to fit the transitions of the /v18=2a n d /v14=1 states. The fit to the population diagram of the main isotopologue includingthe three vibrationally excited states suggests a temperature of ∼250 K (Fig. A.23 ). This explains why our 170 K model needs higher column densities to reproduce the intensities of the /v1 8=2 and /v14=1 transitions. Here again, a more complex model with a nonuniform physical structur e would be necessary to fit all transitions in a consistent way. However, given that our simpleLTE model yields a good fit to the /v1 8=1 transitions of the main isotopologue and all transitions of the13C isotopologues with a single set of parameters, we consider that the derived12C/13C column density ratios are reliable. On the basis of the LTE model obtained above, we were able to identify emission from the doubly substituted13C isotopologue of methyl cyanide,13CH 313CN. One transition at 107 108 MHz is well detected, and a group of transitions at 89 270 MHz is relativel y well detected (Fig. A.24 ). Given that the column density ratio of the singly to doubly substi-tuted isotopologues is very close to the ratio between the main A91, page 6 of 66
10.1051_0004-6361_201527268
page_0005
A. Belloche et al.: Deuterated complex organic molecules in Sgr B2(N2) Table 4. Parameters of our best-fit LTE model (or upper limit) of selected (complex) organic molecules toward Sgr B2(N2). Molecule StatusaNdetbSizecTrotdNeCfΔVgVoffh Nref Ni30 mj (/prime/prime)( K ) ( c m−2)( k m s−1)( k m s−1) CH 3CN, /v18=1⋆d 20 1.40 170 2.2 (18) 1.00 5.4 -0.5 1 y /v18=2 d 8 1.40 170 7.5 (18) 1.00 6.5 –0.5 0.29 y /v14=1 t 1 1.40 170 2.0 (19) 1.00 6.5 –0.5 0.11 y 13CH 3CN, /v1=0 d 8 1.40 170 1.1 (17) 1.18 5.4 –0.5 21 y /v18=1 d 3 1.40 170 1.1 (17) 1.18 5.4 –0.5 21 y CH 313CN, /v1=0 d 7 1.40 170 1.1 (17) 1.18 5.4 –0.5 21 y /v18=1 d 9 1.40 170 1.1 (17) 1.18 5.4 –0.5 21 y 13CH 313CN d 1 1.40 170 4.7 (15) 1.18 5.4 –0.5 466 n CH 3C15N t 0 1.40 170 8.3 (15) 1.18 5.4 –0.5 266 n CH 2DCN d 6 1.40 170 8.3 (15) 1.18 5.4 –0.6 266 n C2H5CN⋆d 154 1.20 150 6.2 (18) 1.38 5.0 –0.8 1 y 13CH 3CH 2CN d 54 1.20 150 1.9 (17) 1.38 5.0 –0.8 32 y CH 313CH 2CN d 38 1.20 150 1.9 (17) 1.38 5.0 –0.8 32 y CH 3CH 213CN d 37 1.20 150 1.9 (17) 1.38 5.0 –0.8 32 y C2H5C15N d 9 1.20 150 1.2 (16) 1.38 5.0 –0.8 500 n CH 2DCH 2CN (out of plane) t 2 1.20 150 3.0 (15) 1.38 5.0 –0.8 2045 n CH 2DCH 2CN (in plane) n 0 1.20 150 <1.5 (15) 1.38 5.0 –0.8 >4091 n CH 3CHDCN t 1 1.20 150 3.0 (15) 1.38 5.0 –0.8 2045 n C2H3CN, /v1=0⋆d 44 1.10 200 4.2 (17) 1.00 6.0 –0.6 1 y /v111=1 d 30 1.10 200 4.2 (17) 1.00 6.0 –0.5 1 y /v115=1 d 20 1.10 200 4.2 (17) 1.00 6.0 –0.5 1 y /v111=2 d 6 1.10 200 4.2 (17) 1.00 6.0 –0.5 1 y 13CH 2CHCN d 10 1.10 200 2.1 (16) 1.38 6.0 –0.6 20 y CH 213CHCN d 9 1.10 200 2.1 (16) 1.38 6.0 –0.6 20 y CH 2CH13CN d 8 1.10 200 2.1 (16) 1.38 6.0 –0.6 20 y C2H3C15N n 0 1.10 200 <3.4 (15) 1.38 6.0 –0.6 >122 n cis-CHDCHCN n 0 1.10 200 <3.4 (15) 1.38 6.0 –0.6 >122 n trans- CHDCHCN n 0 1.10 200 <3.4 (15) 1.38 6.0 –0.6 >122 n CH 2CDCN n 0 1.10 200 <2.1 (15) 1.38 6.0 –0.6 >203 n HC 3N,/v17=1⋆d 6 1.30 170 3.5 (17) 1.44 5.0 –0.7 1 y H13CCCN, /v1=0 d 2 1.30 170 1.7 (16) 1.44 5.0 –0.7 20 y /v17=1 d 4 1.30 170 1.7 (16) 1.44 5.0 –1.0 20 y HC13CCN, /v1=0 d 3 1.30 170 1.7 (16) 1.44 5.0 –0.7 20 y /v17=1 d 3 1.30 170 1.7 (16) 1.44 5.0 –1.0 20 y HCC13CN, /v1=0 d 3 1.30 170 1.7 (16) 1.44 5.0 –0.7 20 y /v17=1 d 3 1.30 170 1.7 (16) 1.44 5.0 –1.0 20 y H13C13CCN t 1 1.30 170 7.2 (14) 1.44 5.0 –0.7 480 n H13CC13CN t 0 1.30 170 7.2 (14) 1.44 5.0 –0.7 480 y HC13C13CN t 1 1.30 170 7.2 (14) 1.44 5.0 –0.7 480 n HC 315N t 0 1.30 170 1.2 (15) 1.44 5.0 –0.7 300 y DC 3N t 0 1.30 170 3.0 (14) 1.51 5.0 –0.5 1144 n CH 3OH, /v1t=1⋆d 16 1.40 160 4.0 (19) 1.00 5.4 –0.2 1 y CH 2DOH t 2 1.40 160 4.8 (16) 1.15 5.4 –0.5 828 n CH 3OD n 0 1.40 160 <2.6 (16) 1.05 5.4 –0.5 >1524 n C2H5OH⋆d 168 1.25 150 2.0 (18) 1.24 5.4 0.0 1 y CH 3CH 2OD n 0 1.25 150 <3.0 (16) 2.96 5.4 0.0 >67 n CH 3CHDOH n 0 1.25 150 <3.0 (16) 2.96 5.4 0.0 >67 n CH 2DCH 2OH (out of plane) n 0 1.25 150 <3.0 (16) 2.96 5.4 0.0 >67 n CH 2DCH 2OH (in plane) n 0 1.25 150 <2.1 (16) 2.96 5.4 0.0 >96 n CH 3OCHO, /v1t=0⋆d 90 1.50 150 1.2 (18) 1.23 4.7 –0.4 1 y /v1t=1 d 35 1.50 150 1.2 (18) 1.23 4.7 –0.4 1 y CH 2DOCHO (out of plane) n 0 1.50 150 <2.5 (16) 1.07 4.7 –0.4 >50 n CH 2DOCHO (in plane) n 0 1.50 150 <7.3 (15) 0.52 4.7 –0.4 >167 n Notes.(a)d: detection, t: tentative detection, n: nondetection.(b)Number of detected lines (conservative estimate, see Sect. 3). One line of a given species may mean a group of transitions of that species that are blended together.(c)Source diameter ( FWHM ).(d)Rotational temperature.(e)Total column density of the molecule. X(Y) means X×10Y.(f)Correction factor that was applied to the column density to account for the contribution of vibrationally or torsionally excited states or other conformers (e.g., gauche for ethanol), in the cases where this contribution was not included in the partition function of the spectroscopic predictions. For deuterated methyl formate, it is the scaling factor used to compute the column densityof each conformer as if it were an independent species. (g)Linewidth ( FWHM ).(h)Velocity offset with respect to the assumed systemic velocity of Sgr B2(N2), Vlsr=74 km s−1.(i)Column density ratio, with Nrefthe column density of the previous reference species marked with a ⋆.(j)Detected (y) or not detected (n) toward Sgr B2(N) (N1 and /or N2) with the IRAM 30 m telescope ( Belloche et al. 2013 ). A91, page 7 of 66
10.1051_0004-6361_201527268
page_0006
A&A 587, A91 (2016) Fig. 1. Deconvolved major, minor, and mean sizes (FWHM) derived for uncontaminated C 2H5CN transitions detected toward Sgr B2(N2) and plotted as a function of upper level energy. The symbols code for the spectral setup (S1 to S5, like in Table 2). In each panel, the dashed line indicates the median value. isotopologue and those that are singly substituted, we consider the identification of13CH 313CN as secure. Our LTE modeling indicates that the15N isotopologue con- tributes significantly to the emission detected at 107 054 MHz and 107 060 MHz (Fig. A.25 ). Since there is no clearly de- tected line, we do not consider this identification as secure. Thederived column density is, therefore, relatively uncertain andshould rather be considered as an upper limit. Finally, using the same parameters as for the other isotopo- logues, we obtain a secure identification of singly deuteratedmethyl cyanide, CH 2DCN, with about six transitions clearly de- tected (Fig. A.26 ). The rotation temperature derived from the population diagram is consistent with the assumed temperaturewithin 2.4σ(Table 3and Fig. A.27 ). The source size can be measured for three of the detected transitions and is found to be consistent with the size derived from the 13C isotopologues. 5.3. Deuterated cyanoacetylene DC 3N Cyanoacetylene is detected toward Sgr B2(N2) in its vibrational ground state (Fig. A.28 ) as well as in many vibrationally excited states: /v17=1( F i g . A.29 ),/v17=2( F i g . A.30 ),/v16=1( F i g . A.31 ), /v15=1a n d /v17=3( F i g . A.32 ), and /v16=/v17=1( F i g . A.33 ). In addition, emission from within the following excited states istentatively detected: /v1 4=1 (one detected line, Fig. A.34 ),/v17=4 and /v15=/v17=1 (significantly contributes to detected signal, but no line individually detected, Fig. A.35 )6,a n d /v16=2 (one detected line, Fig. A.36 ). The three singly substituted13C isotopologues of cyanoacetylene are also clearly detected in their vibrational ground state (Figs. A.37 –A.39 )a n di n /v17=1( F i g s . A.40 –A.42 ). HC13CCN is also detected in /v17=2( F i g . A.43 ) while the two other isotopologues are only tentatively detected in this state(Figs. A.44 andA.45 ). HC 13CCN and HCC13CN are tentatively detected in /v16=1 with one detected line each (Figs. A.46 and 6The current model is somewhat inconsistent with the observed spec- trum at 92 129 MHz (blend of /v17=41 0−2–9 2and 10 4–9 4)a n d 100 431 MHz ( /v15=/v17=1l=0−110–1 0 0) but this is most likely due to resonant interactions between /v17=4a n d /v15=/v17=1, which are not well accounted for in the spectroscopic predictions. The frequen-cies of these transitions may well be o ffb yaf e wM H z( s e eC D M S documentation and Sect. 4.4.33 of Belloche et al. 2013 ).A.47 ). Emission of H13CCCN in /v16=1 significantly contributes to the detected signal, but this state cannot be unambiguously identified (Fig. A.48 ). Two doubly substituted13C isotopologues of cyanoacety- lene, H13C13CCN and HC13C13CN are tentatively detected in their vibrational ground state with one line each (Figs. A.49 andA.50 ). The third, H13CC13CN has no clearly detected line, but the model using the same parameters as the former two iso- topologues is fully consistent with the signal detected around 105 328 MHz (Fig. A.51 ). Therefore we consider this species as tentatively detected too. The15N isotopologue HC 315N is not unambiguously de- tected in its vibrational ground state, but if we assume a 14N/15N isotopic ratio of 300, it contributes significantly to the detected signal at 88 334 MHz and 105 999 MHz and is there- fore included in our model (Fig. A.52 ). The column density of this isotopologue should rather be considered as an upper limit. The fits to the integrated intensity maps suggest that the size of the emission decreases with increasing energy of the vibra- tional state from within which the lines are emitted. Since our model cannot account for a nonunifo rm physical structure, we defined two groups of vibrational states: /v1=0a n d /v17=1 were modeled with a source size of 1 .3/prime/primewhile the higher ex- cited states were modeled assuming 0 .9/prime/prime. The fits to the population diagrams of the singly substituted 13C isotopologues including both /v1=0a n d /v17=1 yield rota- tional temperatures of ∼170–180 K (Table 3,F i g s . A.53 –A.55 ). With a temperature of 170 K and a source size of 1 .3/prime/prime,t h e emission of all isotopologues reported above is well fitted up to/v17=1, except for the vibrational ground state of HC 3N: its transitions are very optically thick ( τmax∼30) and cannot be reproduced with our simple model. For the vibrationally excited states of the main and singly substituted13C isotopologues above /v17=1, we assume a source size of 0 .9/prime/primeand obtain a very good fit to the observed spectra with a temperature of 200 K and a unique column density (divided by 20 for the13C isotopologues) that is 1.5 times higher than for the model of the lower states. Assuming the same parameters as derived above for the vi- brational ground state, we looked for emission of deuterated cyanoacetylene DC 3N. The molecule seems to contribute at a level of∼70% to the signal detected at 101 315 MHz (Fig. A.56 ). The rest of the emission comes from a transition of CH 2CO in its A91, page 8 of 66
10.1051_0004-6361_201527268
page_0007
A. Belloche et al.: Deuterated complex organic molecules in Sgr B2(N2) vibrationally excited state /v19=1. The detection of DC 3N is only tentative, and its column density should rather be considered as an upper limit. 5.4. Deuterated methanol CH 2DOH Methanol and its13Ca n d18O isotopologues are well detected toward Sgr B2(N2). The detected lines and detailed modeling of these species is presente d in a companion paper ( Müller et al. 2016b ). We report in Table 4the parameters derived in that paper for the main isotopologue based on the analysis of allisotopologues. Assuming the same source size and rotational tempera- ture, we obtain a tentative detection of CH 2DOH, with two lines detected at 91 587 MHz (4 1,3e0–40,4e0) and 99 672 MHz (61,5e0–60,6e0), two lines tentatively detected at 85 600 MHz (62,4e1–61,6o1) and 94 563 MHz (1 1,0o1–10,1o1), and a few other lines contributing significantly to the detected signal(Fig. A.57 ). The line appearing at 91 589 MHz in the full syn- thetic model with no counterpart in the observed spectrum corre- sponds to two transitions of acetone (23 18,6–23 17,7of the EE state and 23 18,6–23 17,7of the AE state). While acetone is unambigu- ously detected in our ALMA spectrum of Sgr B2(N2), a signif- icant number of predicted lin es of acetone do not match the ob- served spectrum. The spectrosc opic predictions are not accurate enough for this set of problematic lines, the line at 91 589 MHz being one of those. The ALMA spectrum suggests that the truefrequency could be 91 592 MHz for this acetone line. The source size derived from the maps of the two detected CH 2DOH lines is uncertain but the emission looks compact in the integrated in-tensity maps and is consistent with the source size assumed forthe modeling. 5.5. Upper limits 5.5.1. Deuterated methanol CH 3OD CH 3OD is not unambiguously detected toward Sgr B2(N2). It may significantly contribute to the emission detectedat 90 743 MHz (blend of 10 1,1–92,1and 2 1,0–11,0), 110 951 MHz (41,0–40,0), 111 846 MHz (5 1,0–50,0), and 113 352 MHz (6 1,0– 60,0), but there is no clearly detected line (Fig. A.58 ). Assuming the same parameters as for methanol (Sect. 5.4), we derive a col- umn density upper limit that is a factor 1.8 times lower than the column density tentatively derived for CH 2DOH (Table 4). This upper limit corresponds to the synthetic spectrum shown in red in Fig. A.58 . 5.5.2. Deuterated vinyl cyanide CHDCHCN and CH 2CDCN Many lines of vinyl cyanide are detected in its ground state and vibrationally excited states /v111=1,/v115=1, and /v111=2 (Figs. A.59 –A.62 ). The sizes derived from the corresponding in- tegrated intensity maps tend to decrease with increasing energy, from∼1.2/prime/primeforEup<100 K to∼0.8/prime/primefor higher energy tran- sitions. As a compromise we adopt a source size of 1 .1/prime/prime. With this source size, the analysis of the population diagram yields a temperature of∼200 K (Table 3and Fig. A.63 ). Transitions from within even higher vibrationally excited states are also detectedtoward Sgr B2(N2), but we do not report about these states. Transitions of all three singly substituted 13C isotopologues of vinyl cyanide are also clearly detected (Figs. A.64 –A.66 ;s e e also Müller et al. 2008 for a previous single-dish detection). Only a few lines are su fficiently free of contamination to allowfor a size measurement in the corresponding integrated inten- sity maps. The outcome is more uncertain than for the main iso- topologue, but is consistent with the source size adopted above.Because of the smaller number of detected lines, the population diagrams have a higher dispersion than for the main isotopo- logue and the rotational temperature is less well constrained butthe fits to all three diagrams are consistent with a temperature of about 200 K (Figs. A.67 –A.69 ). As a result of this analysis, we adopt a source size of 1 .1 /prime/prime and a temperature of 200 K for vinyl cyanide and its isotopo- logues. With these parameters, we do not detect the15Ni s o - topologue. We also looked for the singly deuterated species cis- CHDCHCN, trans -CHDCHCN, and CH 2CDCN, but did not de- tect them. Column density upper limits are reported in Table 4. 5.5.3. Deuterated ethanol CH 3CH 2OD, CH 3CHDOH, and CH 2DCH 2OH Ethanol and its13C isotopologues are well detected toward Sgr B2(N2). The detected lines and detailed modeling of these species is presented i n a companion paper ( Müller et al. 2016b ). We report in Table 4the parameters derived in that paper for the main isotopologue based on the analysis of all isotopologues. Assuming the same LTE parameters as for the main isotopo- logue, we searched for all singly deuterated isotopologues ofethanol. None is detected. Upper limits to their column densities are reported in Table 4. 5.5.4. Deuterated methyl formate CH 2DOCHO Methyl formate is clearly seen toward Sgr B2(N2), with dozens of transitions detected in both its ground and first torsionalstates (Figs. A.70 andA.71 ). We derive a median source size of 1.5 /prime/primefrom fits to the integrated intensity maps of its numer- ous uncontaminated lines. The formal fit to its population di-agram including both states yields a rotational temperature of ∼140 K (Fig. A.72 and Table 3). We used a temperature of 150 K in our model, which fits the ALMA spectrum very well,apart from a few discrepancies that we describe now. The rea- son why the synthetic spectrum of the ground state poorly fits the ALMA spectrum at 100 080 MHz is unclear. It may be dueto the nearby HC 3N 11–10 transition at 100 076 MHz, which is probably affected by self-absorption and /or spatial filtering and is by far overestimated by our simple LTE model. Thediscrepancy around 110 226 MHz is due to contamination by dif- fuse cloud absorption in 13CO 1–0 that is not yet included in our full model. Similar contamination by c-C3H2absorption features not yet implemented in our full model likely explains the smalldiscrepancies for the /v1 t=1 transitions around 85 370 MHz. Assuming the same LTE parameters as for the main isotopo- logue, we searched for the in-plane and out-of-plane conformersof CH 2DOCHO toward Sgr B2(N2) but none of them is detected. Upper limits to their individual column densities are reported in Table 4. 6. Discussion 6.1. Comparison to other observations The levels of deuterium fractionation derived in Sect. 5for (com- plex) organic molecules toward Sgr B2(N2) are summarized inTable 5and shown in Fig. 2. A91, page 9 of 66
10.1051_0004-6361_201527268
page_0008
A&A 587, A91 (2016) Table 5. Deuterium fractionation of selected (complex) organic molecules toward Sgr B2(N2) compared to predictions of astrochem-ical models. Molecule Statusa[XD]/[XH] N2bT14cA12d %% % CH 2DCN d 0.38 3.6–0.15 2.3–6.3 CH 2DCH 2CN (oop) t 0.05 – – CH 2DCH 2CN (ip) n <0.024 – – CH 3CHDCN t 0.05 – – cis-CHDCHCN n <0.8 – – trans- CHDCHCN n <0.8 – – CH 2CDCN n <0.5 – – DC 3N t 0.09 – 5.4–1.1 CH 2DOH t 0.12 5.5–0.51 2.4–2.4e CH 3OD n <0.07 3.5–0.3 –e CH 3CH 2OD n <1.5 – – CH 3CHDOH n <1.5 – – CH 2DCH 2OH (oop) n <1.5 – – CH 2DCH 2OH (ip) n <1.0 – – CH 2DOCHO (oop) n <2.0 14–0.43f– CH 2DOCHO (ip) n <0.6 7–0.22f– Notes. The notations oop and ip describe the position of the deuterium and stand for out of plane and in plane, respectively.(a)d: detection, t: tentative detection, n: nondetection.(b)Deuterium fractionation mea- s u r e dt o w a r dS g rB 2 ( N 2 ) .(c)Deuterium fractionation predicted by the model of Taquet et al. (2014 ) in the hot corino at the beginning and end of the Class 0 phase.(d)Deuterium fractionation predicted by the model ofAikawa et al. (2012 ) in the hot corino at the beginning and end of the Class 0 phase.(e)The model of Aikawa et al. (2012 ) was not de- signed to predict the abundance ratios of deuterated isomers: it assumesstatistical branching ratios. (f)The model of T a q u e te ta l . (2014 ) does not distinguish between the in-plane and out-of-plane conformers. Thevalues listed here assume a statistical distribution (2:1). 6.1.1. Deuterated methyl cyanide The detection of CH 2DCN toward Sgr B2(N2) is the most se- cure among the deuterated species reported here (Sect. 5.2and Fig.A.26 ). On the basis of the LTE modeling of methyl cyanide and its various isotopologues, we derive a deuterium fraction-ation of 0.4% for this molecule. This is a factor 2.6 lower than the fractionation reported by Gerin et al. (1992 )t o w a r d Orion KL (1%). Along with this first interstellar detection, theseauthors also reported a tentative detection toward the hot core G34.26+0.15 that, if true, would indicate a similar level of deu- terium fractionation as toward Orion KL. The di fference with the level measured in Sgr B2(N2) is probably not significant because the Orion KL and G34.26 +0.15 values may su ffer, as mentioned by these authors, from a lack of knowledge of the source sizeand opacity of the lines of the main isotopologue. A detection of deuterated methyl cyanide toward the Class 0 (low-mass) protostar IRAS 16293–2422 was also reported in Taquet et al. (2014 ) based on an unpublished analysis. They quote a deuterium fractionation of 1 .3%, a factor 3.4 higher than the one obtained for Sgr B2(N2). 6.1.2. Deuterated ethyl cyanide Both deuterated isotopologues of ethyl cyanide are tenta-tively detected toward Sgr B2(N2) (Sect. 5.1and Figs. A.10 andA.11 ). We derive a deuterium fractionation of ∼0.05% for both CH 2DCH 2CN (in its out-of-plane conformation) and the chiral molecule CH 3CHDCN. The upper limit obtained for theFig. 2. Deuterium fractionation of (complex) organic molecules toward Sgr B2(N2). Secure detections are indicated with a filled square, ten-tative detections with an empty square, and upper limits with an arrowpointing to the left. The notations oop and ip describe the position ofthe deuterium and stand for out of plane and in plane, respectively. in-plane conformer of CH 2DCH 2CN (Fig. A.12 ) is still consis- tent with the expectation that it should be half as abundant asthe out-of-plane conformer. If we assume this expected ratio,then the total deuterium fractionation for CH 2DCH 2CN would be∼0.075%, a factor 1.5 times higher than for CH 3CHDCN. This would be consistent with th e statistical expectation because the methyl group at the end of the carbon chain has three equiva- lent hydrogen atoms while the middle chain group has only two. The deuterium fractionation derived for ethyl cyanide to- ward Sgr B2(N2) is nearly one order of magnitude lower thanfor methyl cyanide (0.4%), but similar within a factor two to methanol (0.12%) and cy anoacetylene (0.09%). Margulès et al. (2009 ) reported a detection of the 15Ni s o - topologue of ethyl cyanide toward Orion KL but obtained onlyan upper limit for CH 2DCH 2CN. They derived a column density ratio [CH 2DCH 2CN (oop)]/[C2H5C15N]<0.33, which trans- lates into [CH 2DCH 2CN (oop)]/[C2H5CN]<0.2% using the 14N/15N isotopic ratio of 148 ±74 derived by Daly et al. (2013 ). Daly et al. (2013 ) claimed tentative detections of both deuter- ated isotopologues of ethyl cyanide with a deuterium fractiona-tion of 2% based on the same survey of Orion KL. No detected transitions are shown in that study, though, and their Table 3 ac- tually reports upper limits for the deuterated species. Given thelower deuterium fractionation obtained by Gerin et al. (1992 ) for methyl cyanide toward Orion KL (1%) and the order of magnitude difference in deuterium fractionation between methyl cyanide and ethyl cyanide obtained here toward Sgr B2(N2), a deuterium fractionation of 2% for ethyl cyanide in Orion KL sounds unlikely and questions the tentative detection of Daly et al. (2013 ). 6.1.3. Deuterated methanol CH 2DOH is tentatively detected toward Sgr B2(N2) (Sect. 5.4 and Fig. A.57 ). The deuterium fractionation we derive for this isotopologue is 0.12%, a factor ∼3 lower than the value we A91, page 10 of 66
10.1051_0004-6361_201527268
page_0009
A. Belloche et al.: Deuterated complex organic molecules in Sgr B2(N2) obtain for methyl cyanide, the d euterated species of which is securely identified in our ALMA spectrum of Sgr B2(N2) (Sect. 6.1.1 ). Therefore, even if the di fference in deuterium fractionation between methanol and methyl cyanide is a priori surprising, the fact that it is lower for methanol gives us more confidence in the detection of CH 2DOH. The deuterium frac- tionation derived for CH 2DOH is a factor five lower than to- ward the Compact Ridge in Orion KL (0.58%, Neill et al. 2013 ), about one order of magnitude lower than toward the high-massprotostellar objects surveyed by Fontani et al. (2015 )a n dt h e intermediate-mass protostar NGC 7129 FIRS 2 (2%, Fuente et al. 2014 ), and more than two orders of magnitude lower than toward Class 0 protostars which have values between 19%and 33% ( Parise et al. 2006 ) 7. Evidence for the presence of CH 3OD is too tenuous in our ALMA spectrum of Sgr B2(N2) to claim a detection, even a tentative one (Sect. 5.5.1 ). Still, we cannot completely exclude that CH 3OD is present at the level indicated by our upper limit. The synthetic spectrum shown in Fig. A.58 indicates that a large fraction (>50%) of the flux density detected at 90 744, 99 964, 110 951, and 113 352 MHz may well be emitted by this molecule. If this is true, the deuterium fractionation of methanolwould then be∼0.07% for CH 3OD. This would be nearly one order of magnitude lower than toward the Compact Ridge in Orion KL (0.5%, Neill et al. 2013 ) and about 50 times lower than toward Class 0 protostars (1.6%–4.7%, Parise et al. 2006 ). A detection of CH 3OD toward Sgr B2 was reported by Gottlieb et al. (1979 ) with the 36 foot radio telescope of the National Radio Astronomy Observatory at Kitt Peak ( HPBW∼ 74/prime/prime). These authors detected a line at the frequency ex- pected for the pair of partially blended transitions 2 −1–1−1E (90 703.6 MHz, Eu/kB=11.3K )a n d2 0–10A (90 705.8 MHz, Eu/kB=6.5 K), but they did not detect the nearby 2 1–11Et r a n - sition (90 743.5 MHz, Eu/kB=15.6K ) ,w h i c hi si nf a c te x - pected to be partially blended with the 10 1–92A transition (90 741.7 MHz, Eu/kB=124 K). The former two transitions are blended with deep HNC absorption features produced bydiffuse clouds along the line of sight in our ALMA spectrum of Sgr B2(N2). This prevents th eir detection in our spectrum (see Fig. A.58 ). Our LTE model shows that, for a temperature of 160 K, the latter two transitions are expected to be as strong as the former two. We conclude from this that either the assignment of the 90 704 MHz line to CH 3OD in the Kitt Peak spectrum was not correct, or the line reported by Gottlieb et al. (1979 ) traces low-excitation emission of CH 3OD. Given that the line detected in emission in the Kitt Peak sp ectrum dominates over the ab- sorption features, opposite to what is seen in the ALMA spec-trum, this emission line, if real, must come from a region more extended than the Sgr B2 continuum emission that is absorbed by the diffuse clouds along the line of sight. Such an extended emission would be filtered out in our ALMA spectrum. Gottlieb et al. (1979 ) derived a ratio [CH 3OD]/[13CH 3OH]∼0.18 for Sgr B2. This translates into [CH 3OD]/[CH 3OH]∼0.7%, assuming a12C/13C isotopic ratio of 25 as derived for methanol toward Sgr B2(N2) in the companion paper Müller et al. (2016b ). This is an order of 7There was an issue with the spectroscopic predictions used in the early studies reporting CH 2DOH column densities (B. Parise, priv. comm.). We compared the Sμ2values listed in Table 1 of Parise et al. (2002 ) ,w h i c hw e r ea l s ou s e di n Parise et al. (2006 ), with the cur- rent JPL catalog. The new values of the selected transitions are a factor2.1±0.4 times higher on average than the old values. The partition func- tion is the same in both cases. As a result, the column densities reportedfor CH 2DOH in both articles were overestimated by a factor of ∼2.magnitude higher than our upper limit of 0.07% derived in Sect. 5.4. This discrepancy seriously questions the detection of CH 3OD reported by Gottlieb et al. (1979 )t o w a r dS g rB 2 , unless deuteration of methanol is more e fficient by one order of magnitude on large scales in the Sgr B2 cloud compared to the embedded hot cores. Comito et al. (2003 ) derived an abundance ratio [HDO]/[H2O]∼0.06% toward the Sgr B2 hot cores and their T<100 K envelope, and even lower values of ∼0.013% and∼0.02% (uncertain within a factor two) were obtained for [DCN]/[HCN] and [DCO+]/[HCO+] in the molecular ridge close to Sgr B2(M) ( Jacq et al. 1999 ). Deuterium fractionation thus does not appear to be generally more e fficient on larger scales in Sgr B2, which again questions the detection of CH 3OD reported by Gottlieb et al. (1979 ). 6.1.4. Deuterated cyanoacetylene The detection of DC 3N reported toward Sgr B2(N2) is only ten- tative (Sect. 5.3and Fig. A.56 ). We obtain a deuterium fraction- ation of 0.09%, similar to the values obtained for methanol and ethyl cyanide. DC 3N was first detected toward TMC 1 with a deuterium fractionation of 2–8% ( Langer et al. 1980 ), revised to a lower value of 1.5% by Turner (2001 ). High values were reported with single dish telescopes for a number of other cold dense cores (5%–10%, Howe et al. 1994 )a sw e l la sf o rap r o t o - star in a stage of “Warm Carbon-Chain Chemistry” ( ∼3%,Sakai et al. 2009 ). A tentative detection toward the Compact Ridge and the Hot Core of Orion KL was recently reported with a deu-terium fractionation of 1 .5%±0.9% ( Esplugues et al. 2013 ). A tentative detection was also recently reported toward the high- mass protostar NGC 2264 CMM3 (1.8% ±1.5%, Watanabe et al. 2015 ). The deuterium fractiona tion of cyanoacetylene ten- tatively derived toward Sgr B2(N2) is thus at least one order of magnitude lower than in Orion and NGC 2264 CMM3 (if con-firmed) and even two orders of magnitude lower than in cold dense gas. 6.1.5. Deuterated vinyl cyanide The column density upper limits reported in Sect. 5.5.2 yield deuterium fractionations <0.8%,<0.8%, and<0.5% for cis-CHDCHCN, trans- CHDCHCN, and CH 2CDCN, respectively. We are not aware of any reliable detection of deuterated vinyl cyanide in the interstellar medium. 6.1.6. Deuterated ethanol The column density upper limits reported in Sect. 5.5.3 yield deuterium fractionations <1.5%,<1.5%,<1.5%, and<1.0% for CH 3CH 2OD, CH 3CHDOH, and the out-of-plane and in- plane conformers of CH 2DCH 2OH, respectively. The latter two translate into a total deuterium fractionation <2.5% for CH 2DCH 2OH. These upper limits are about one order of mag- nitude higher than the deuterium fractionation measured for methanol (Sect. 6.1.3 ). They are thus not very constraining. We are not aware of any detection of deuterated ethanol in theinterstellar medium. 6.1.7. Deuterated methyl formate The column density upper limits reported in Sect. 5.5.4 yield deuterium fractionations <2.0% and<0.6% for the out-of-plane and in-plane conformers of CH 2DOCHO, respectively. This A91, page 11 of 66
10.1051_0004-6361_201527268
page_0010
174901-3 P.-M. Lam and Y. Zhen J. Chem. Phys. 143, 174901 (2015) g(f)+1 2Asin4α r2+U(r)−sin 2α 2rM=0. (10) Three equations (8)–(10) can be solved to obtain the remaining three unknowns quantities α,r, and M. In the high-force limit, Neukirch and Marko28have given the solutions in closed form. Define a quantity Kas K= 9π/8ν2LBkBT/g(f). (11) The plectoneme radius rand angleαare given by r=logK/(2κD), (12) α=2r2g(f)/(3A)1/4. (13) The torque Mis given by M=2A rsin3αcosα cos 2α. (14) A Taylor expansion of this expression for small αand substituting the results, Eqs. (12) and (13), for αandryields M≈[(32/27)A]1/4g(f)3/4/√κD × log 9π/8ν2LBkBT/g(f)1+α2.(15) Another quantity of experimental interest is the slope of the average extension q=∂⟨X⟩/∂∆Lk=−∂2G/∂f∂∆Lk =−2π∂M/∂f. Using Eq. (10) for M, this becomes q =−4πrg′(r)/sin 2α. Taylor expanding this for small αand substituting Eqs. (12) and (13) for αandryields q=(6A κD2g(f))1/4 log 9π/8ν2LBkBT/g(f) ×g′(f)(1+2α2/3). (16) Using Eqs. (15) and (16) and g(f)=f−kBT f/A, the slope and torque calculated are in qualitative agreement with experiment.28Experimental values of A/kBT=46, 47, 44, 45 nm at 50, 100, 200, and 500 nM salt and C/kBT=94 nm are used in the calculation. We will show in Sec. III that using a more accurate form of g(f)can significantly improve on the agreement with experiment. III. CALCULATION USING AN IMPROVED FREE ENERGY In this section, we give our calculation of the slope and the torque using an improved form of the untwisted free energy. The force-extension curve in the worm like chain (WLC) model is given by the widely used interpolation formula26 f=(kBT) LpX L+1 4( 1−X L)−2 −1 4, (17) where Lphere is the persistence length, related to the bending rigidity A, byA=kBT Lp, and Xis the extension. The negative of the free energy per unit length g(f)is obtained by a Legendre transform Lg(f)=f X−W(X), (18)where W(X)=X 0dX′f(X′) (19) is the work done in extending the polymer. The functions gandWdepend also on the persistence length Lp. From Eq. (17), the extension Xis an implicit function of the force f. Since the extension is a single-valued, monotonic increasing function of f, we can define the inverse function Xf(f)which gives the extension Xas a function of the force f. Even though this function cannot be obtained analytically, it can be calculated numerically to high accuracy. Substituting Eq. (17) into Eq. (19), the function Wcan be calculated analytically, W(X(f)) =LkBT 4LpXf(f) L( 2Xf(f) L−1) +( 1−Xf(f) L)−1.(20) From Eq. (18), the negative of the free energy per unit length is given as a function of the force fby g(f)=1 Lf Xf(f)−kBT 4Lp ×Xf(f) L( 2Xf(f) L−1) +( 1−Xf(f) L)−1. (21) We will use this form of the free energy in Eqs. (15) and (16) to calculate the torque Mand slope q. From Eq. (16), in order to calculate the slope q, the derivative of gwith respect to fis needed. From Eqs. (18) and (19), this is given by Lg′(f)=Xf(f). (22) In Fig. 2, we show our calculation of the slope of the average extension q=∂⟨X⟩/∂∆Lkobtained using Eq. (16), FIG. 2. Comparison of experimental and theoretical slopes q=∂⟨X⟩/∂∆Lk of the average extension, as a function of the applied force, for 50, 100, 200, and 500 mM salt (top to bottom). Circles are experimental data. Full lines are our theoretical results using a better form of the free energy. Dashed lines are theoretical results using approximate form of the free energy.
10.1063_1.4934988
page_0003
174901-4 P.-M. Lam and Y. Zhen J. Chem. Phys. 143, 174901 (2015) FIG. 3. Comparison of experimental and theoretical torque as a function of the applied force, for 50, 100, 200, and 500 mM salt (top to bottom). Circles are experimental data. Full lines are our theoretical results using a better form of the free energy. Dashed lines are theoretical results using approximate form of the free energy. withgandg′given by Eqs. (21) and (22), together with results obtained using the approximate forms for gandg′. The experimental data are directly taken from Fig. 2 of Ref. 28. The data in Ref. 28 are obtained from Ref. 21. The slope in Ref. 21 is a dimensionless quantity defined as ˜ q=L−1∂⟨X⟩/∂σ, with σ=(Lk−Lk0)/Lk0, where Lk0≈1500 is the linking number of the DNA molecule under no external tension or torque. Our slope qis related to ˜ qbyq=(L/Lk0)˜q=(5.4µm/1500 )˜q =(3.6 nm )˜q. The experimental data given in Ref. 28 are actually a factor (−3.6 nm )times the data in Ref. 21. In Figure 3, we present the results of our calculation for the torque M, using Eqs. (15) and (21), together with results obtained using the approximate form, compared with the experimental data, taken from Fig. 3 of Ref. 21. We can see that this better form of the free energy improves significantly the agreement with experiment. The agreement with experiment is now surprisingly good, except for low salt concentrations. IV. CONCLUSION We have shown that by using a better form of the free energy for the stretched but untwisted part of the DNA, the Neukirch-Marko model can give quantitative agreement with experimental results. There is still some disagreement at low salt concentration, but this is probably due to the inadequacy of the Debye–Hückel approximation of the Poisson-Boltzmann equation, which results in imperfect screening of the elec- trostatic potential at these low salt concentrations. It was mentioned in Ref. 28 that the disagreement with experiment may be due to the neglect of confinement entropy.36Since our results using a better free energy already yield quantitative agreement with experiment, the e ffect of confinement entropy is probably small. Our calculation is based on the model of Neukirch and Marko.28This theory is an analytic theory, with analyticexpressions for the slope and torque as functions of the tension. In order to arrive at this theory, several reasonable simplications have been introduced. It does not incorporate thermal fluctuations in plectoneme. The argument is that at least at higher tensions, the fluctuations are small and can as a consequence be neglected. It also neglects multi- plectoneme e ffects. The use of a two-cylinder repulsion in the Debye–Hückel regime is a rough approach not taking into account the e ffect on plectoneme angle as was shown to be important by Ubbink and Odijk.35More recent models37,38 have taken these e ffects into account. In Ref. 37, the authors give results of the slope versus tension, in very good agreement with experiment. However, for this quantity, the original theory of Neukirch and Marko also gives good agreement with experiment. It is the torque versus tension results in the Neukirch-Marko theory that show the largest disagreement with experiment, especially for low tension and low salt concentrations. For the torque versus tension result, the result of Ref. 37 is not so good. In more recent work,38Marko and Neukirch have also incorporated the above mentioned e ffects in their model, but unfortunately they do not give any new torque versus tension results. Notwithstanding the clearly better agreement between theory and experiment achieved in this work, one notes, how- ever, that it holds as far as the Debye–Hückel approximation of the Poisson-Boltzmann theory remains valid, i.e., for high screening /salt concentration only. As one can see from Figs. 2 and 3, the agreement with experiment deteriorates at low salt concentration for both the slope and the torque. A closer inspection of the q-f variation, shown in Fig. 2, indicates that the agreement with experiment at higher applied tensions (when f >3 pN at 500 mM and f >1 pN at 200 mM). This is puzzling because the expressions for the free energy g(f)and twist modulus Cs(f)should be correct for large fand the fluctuations in plectoneme and multi-plectoneme e ffects neglected in the model should also decrease with tension. Dhar and Chaudhuri39and Samuel and Sinha40have explored e ffects that go beyond the high force limit ( g(f)=f−kBT f/A). At these high forces, such e ffects may be relevant. It should also be pointed out that the force-extension formula given in Eq. (17) is only an interpretation formula which is convenient for calculation and should not be considered as a substitute for analytical or semi-analytic theoretical models in Refs. 39 and 40. In particular, the force-extension curve given by Eq. (19) did not take into account the entropy of the chain, even when no external force is applied, as pointed out by Neumann.41Also, in this high tension limit, one would have to include the e ffects of thermal fluctuations on DNA elasticity, as studied by Kulic et al.42,43and Sinha and Samuel.40,44Finally, the Legendre transform (Eqs. (18) and (19)) was used because we are considering the long DNA limit. If one were to look at shorter chains, with chain lengths comparable to the persistence length of 50 nm, one would need to work with Laplace transforms instead.41,45,46This is because such short chains are not in the thermodynamic limit and one has to distinguish between the isometric ensemble in which the chain ends are held fixed and the applied force is allowed to fluctuate and the isotensional ensemble in which the applied force is held fixed and the chain lengths are allowed to fluctuate. Only in the thermodynamic
10.1063_1.4934988
page_0004
174901-5 P.-M. Lam and Y. Zhen J. Chem. Phys. 143, 174901 (2015) limit in which the chain lengths are infinitely long, do the two ensembles yield identical results. 1D. Branzei and M. Foiani, Nat. Rev. Mol. Cell Biol. 11, 208 (2010). 2N. R. Cozzarelli, G. J. Cost, M. Nollmann, T. Viard, and J. E. Stray, Nat. Rev. Mol. Cell Biol. 7, 580 (2006). 3F. Kouzine and D. Levens, Front. Biosci. 12, 4409 (2007). 4D. Kowalski and M. J. Eddy, EMBO J. 8, 4335 (1989). 5L. F. Liu and J. C. Wang, Proc. Natl. Acad. Sci. U. S. A. 84, 7024 (1987). 6L. Postow, C. Ullsperger, R. W. Keller, C. Bustamante, A. V . V ologodskii, and N. R. Cozzarelli, J. Biol. Chem. 276, 2790 (2001). 7A. Revyakin, R. H. Ebright, and T. R. Strick, Proc. Natl. Acad. Sci. U. S. A. 101, 4776 (2004). 8G. Lia, D. Bensimon, V . Croquette, J.-F. Allemand, D. Dunlap, D. Lewis, S. Adhya, and L. Finzi, Proc. Natl. Acad. Sci. U. S. A. 100, 11373 (2003). 9N. Cozzarelli, Science 207, 953 (1980). 10T. Strick, V . Croquette, and D. Bensimon, Nature 404, 901 (2000). 11T. Strick, J.-F. Allemand, D. Bensimon, A. Bensimon, and V . Croquette, Science 271, 1835 (1996). 12T. R. Strick, J.-F. Allemand, D. Bensimon, and V . Croquette, Biophys. J. 74, 2016 (1998). 13T. R. Strick, J.-F. Allemand, D. Bensimon, R. Laverhy, and V . Croquette, Physica A 263, 392 (1999). 14T. Strick, J.-F. Allemand, V . Croquette, and D. Bensimon, Prog. Biophys. Mol. Biol. 74, 115 (2000). 15D. A. Koster, V . Croquette, C. Dekker, S. Shuman, and N. H. Dekker, Nature 434, 671 (2015). 16B. Taneja, B. Schnurr, A. Slesarev, J. F. Marko, and A. Mondragon, Proc. Natl. Acad. Sci. U. S. A. 434, 14670 (2007). 17O. A. Saleh, S. Bigot, F.-X. Barre, and J.-F. Allemand, Nat. Struct. Mol. Biol. 12, 436 (2005). 18R. Seidel, G. G. P. Bloom, C. Dekker, and M. D. Szczelkun, EMBO J. 27, 1388 (2008). 19G. Lia, E. Praly, H. Ferreira, C. Stockdale, Y . C. Tse-Dinh, D. Dunlap, V . Croquette, D. Bensimon, and T. Owen-Huges, Mol. Cell 21, 417 (2006).20Z. Bryant, M. D. Stone, J. Gore, S. B. Smith, N. R. Cozzarelli, and C. Bustamante, Nature 424, 338 (2003). 21F. Mosconi, J.-F. Allemand, D. Bensimon, and V . Croquette, Phys. Rev. Lett. 102, 078301 (2009). 22S. Smith, L. Finzi, and C. Bustamante, Science 258, 1122 (1992). 23C. Bustamante, J. F. Marko, S. Smith, and E. D. Siggia, Science 265, 1599 (1994). 24A. V . V ologodskii, Macromolecules 27, 5623 (1994). 25T. Odijk, Macromolecules 28, 7016 (1995). 26J. F. Marko and E. D. Siggia, Macromolecules 28, 8759 (1995). 27J. F. Marko, Phys. Rev. E 76, 021926 (2007). 28S. Neukirch and J. F. Marko, Phys. Rev. Lett. 106, 138104 (2011). 29N. Clauvelin, B. Audoly, and S. Neukirch, Macromolecules 41, 4479 (2008). 30J. F. Marko, in Mathematics of DNA Structure, Function and Interactions , edited by C. J. Benham et al. , The IMA V olumes in Mathematics and it Applications V ol. 150 (Springer Science +Business Media, 2009). 31J. F. Marko and E. D. Siggia, Science 265, 506 (1994). 32J. D. Moroz and P. Nelson, Proc. Natl. Acad. Sci. U. S. A. 94, 14418 (1997). 33D. Stigter, J. Colloid Interface Sci. 53, 296 (1975). 34J. A. Schellman and D. Stigter, Biopolymers 16, 1415 (1977). 35J. Ubbink and T. Odijk, Biophys. J. 76, 2502 (1999). 36J. F. Marko and E. D. Siggia, Phys. Rev. E 52, 2912 (1992). 37M. Emanuel, G. Lanzani, and H. Schiessel, Phys. Rev. E 88, 022706 (2013). 38J. Marko and S. Neukirch, Phys. Rev. E 85, 011908 (2012). 39A. Dhar and D. Chaudhuri, Phys. Rev. Lett. 89, 065502 (2002). 40J. Samuel and S. Sinha, J. Phys.: Condens. Matter 25, 465102 (2013). 41R. Neumann, Biophys. J. 85, 3418 (2003). 42I. M. Kulic, H. Mohrbach, V . Lobaskin, R. Thaokar, and H. Schiessel, Phys. Rev. E 72, 041905 (2005). 43I. M. Kulic, H. Mohrbach, V . Lobaskin, R. Thaokar, and H. Schiessel, Phys. Rev. E 75, 011913 (2007). 44S. Sinha and J. Samuel, Phys. Rev. E 85, 041802 (2012). 45D. Keller, D. Swigon, and C. Bustamante, Biophys. J. 84, 733 (2003). 46S. Sinha and J. Samuel, Phys. Rev. E 71, 021104 (2005).
10.1063_1.4934988
page_0005
J. Chem. Phys. 145, 114901 (2016); https://doi.org/10.1063/1.4962516 145, 114901 © 2016 Author(s).On the influence of the intermolecular potential on the wetting properties of water on silica surfaces Cite as: J. Chem. Phys. 145, 114901 (2016); https://doi.org/10.1063/1.4962516 Submitted: 17 June 2016 • Accepted: 29 August 2016 • Published Online: 16 September 2016 E. Pafong , J. Geske and B. Drossel ARTICLES YOU MAY BE INTERESTED IN A reactive molecular dynamics simulation of the silica-water interface The Journal of Chemical Physics 132, 174704 (2010); https://doi.org/10.1063/1.3407433 Contact angles from Young’s equation in molecular dynamics simulations The Journal of Chemical Physics 147, 084708 (2017); https://doi.org/10.1063/1.4994088 TRAVIS—A free analyzer for trajectories from molecular simulation The Journal of Chemical Physics 152, 164105 (2020); https://doi.org/10.1063/5.0005078
10.1063_1.4962516
page_0000
THE JOURNAL OF CHEMICAL PHYSICS 145, 114901 (2016) On the influence of the intermolecular potential on the wetting properties of water on silica surfaces E. Pafong,a)J. Geske, and B. Drossel Institut für Festkörperphysik, Technische Universität Darmstadt, Hochschulstr. 6, 64289 Darmstadt, Germany (Received 17 June 2016; accepted 29 August 2016; published online 16 September 2016) We study the wetting properties of water on silica surfaces using molecular dynamics (MD) simu- lations. To describe the intermolecular interaction between water and silica atoms, two types of interaction potential models are used: the standard BródkA and Zerda (BZ) model and the Gulmen and Thompson (GT) model. We perform an in-depth analysis of the influence of the choice of the potential on the arrangement of the water molecules in partially filled pores and on top of silica slabs. We find that at moderate pore filling ratios, the GT silica surface is completely wetted by water molecules, which agrees well with experimental findings, while the commonly used BZ surface is less hydrophilic and is only partially wetted. We interpret our simulation results using an analytical calculation of the phase diagram of water in partially filled pores. Moreover, an evaluation of the contact angle of the water droplet on top of the silica slab reveals that the interaction becomes more hydrophilic with increasing slab thickness and saturates around 2.5–3 nm, in agreement with the experimentally found value. Our analysis also shows that the hydroa ffinity of the surface is mainly determined by the electrostatic interaction, but the van der Waals interaction nevertheless is strong enough that it can turn a hydrophobic surface into a hydrophilic surface. Published by AIP Publishing. [http: //dx.doi.org /10.1063 /1.4962516] I. INTRODUCTION Water is an essential material in our everyday life and is the most used solvent for chemical and biological reactions. Water molecules are highly polar, forming hydrogen-bonded networks and sharing hydrogen bonds with other molecules. In particular, water confined within nanoscale geometries of hydrophilic surfaces is subject to two competing interactions: the hydrophilic interactions between water molecules and those between water and surface molecules. One of the standard systems for studying hydrophilic interactions with water is silica nanopores such as sols- gels,1mesoporous silica (MCM-41),2–7Vycor-glasses,8–10 controlled pore glasses (CPGs).11–15Water confined in silica nanopores or near silica flat surfaces is a topic which has attracted considerable attention,16–22mainly because of the relevance of the water-silica interaction in understanding the water transport in porous rocks,23 nanofluidic devices,24heterogeneous catalysis in mesoporous materials,15,25and permeation through membrane channels.26 Experimental investigations of water in silica nanopores have been carried out using NMR spectroscopy,2,12X-ray and neutron di ffraction,5,6,9,10,27,28quasi-elastic neutron scattering,3,4Small-angle neutron scattering (SANS),8and optical Kerr-e ffect spectroscopy,1showing that the dynamics of water in such pores is slow in comparison to the dynamics of bulk water. This originates from the strong binding or trapping of water molecules by silica surfaces as found by experimental measurements conducted with MCM-41 as well as CPG pores2,12resulting in a complete coverage of the pore a)priscel@fkp.tu-darmstadt.desurface at even moderate hydration levels. Accordingly, the intermolecular interactions between water and silica surfaces in MD simulations should be set up such that the experimental results are reproduced. The influence of the filling ratios on the wetting properties of water in silica nanopores has been studied by previous MD research.29–32Such investigations are motivated by the fact that in experiments the fluid is placed on top of a porous surface and flows to enter the pores. In the mentioned MD simulations, it was demonstrated that at all hydration levels water molecules are absorbed by the Vycor material. However, they have not shown to what extent the pore surface is wetted, whether it is only partially wetted or rather completely wetted (as expected from measurements2,12). Moreover, the confinement near such hydrophilic surface was found to substantially alter the dynamic behaviour of water, depending on the filling ratio,19,33but it has not been checked how this relates to the configurations that water can take inside the pore. In all these previous MD analyses, the (12-6) Lennard-Jones (LJ) potential and the partial charges assigned to each silica atom site are chosen according to BródkA and Zerda (BZ).34In this model, the LJ potential parameters for silica oxygen atoms are approximated from the Kirkwood-Mueller formula,35while no LJ interaction centers are assigned to silicon and hydrogen (of the silanol groups) as they are small in size and possess a low polarizability. In the present investigation, we have found that water molecules do not completely wet the BZ silica model surface at intermediate hydration levels. For this reason, the silica model recently introduced by Gulmen and Thompson36(GT) has been tested. The GT36potential is defined similarly to the silica potential by BZ,34however, a weak short-ranged interaction for silicon and hydrogen 0021-9606/2016/145(11)/114901/9/$30.00 145, 114901-1 Published by AIP Publishing.
10.1063_1.4962516
page_0001
114901-2 Pafong, Geske, and Drossel J. Chem. Phys. 145, 114901 (2016) atoms has been added and the partial charges on each silica atom are increased. We probed the performance of both silica models by analyzing the wetting behaviour of water on silica surfaces and comparing to the experimental results.2,12,37–39 In the following, we present the results of MD simulations of water in a cylindrical silica nanopore of roughly 4 nm diameter and 6 .1 nm height. Additionally, MD simulations of water droplets wetting silica slabs of varying thicknesses were performed. The silica nanopores and slabs were created in our group, with the silanols molecules uniformly distributed on the surface. For water in the silica pore, we evaluate the minimum number of water layers necessary to completely wet a silica surface by looking at the radial and angular density distribution, as well as the number of hydrogen bonds formed for di fferent filling ratios of water in the pore. Furthermore, a phase diagram of the surface tension of the different configurations adopted by water molecules in the nanopore is calculated analytically, providing deeper insights into the relation between the interaction energies and the water arrangement in the pore. To complete the work, the contact angle of a water droplet on a silica flat slab is evaluated and compared between the two model surfaces. Previous MD investigations have used the contact angle investigation to approximate LJ parameters between water and silica atoms40,41but they have not stated clearly whether the simulations were performed in such a way that the periodic images provided by the periodic boundary conditions do not influence the contact angle evaluated. In our investigation, we run the simulations without periodic boundary conditions to avoid this issue. Previous experimental results showed that the wetting properties do not only involve atoms of layers in the vicinity of the interface but also the atoms located deeply inside the slab material.37–39,42 Therefore, we measure how the contact angle changes with the thickness of the slab, showing that a thickness of 2.5–3 nm is sufficient for MD simulations. In order to disentangle the contributions of the interfacial electrostatic and van der Waals (VdW) interactions on the contact angle, we varied these two contributions in our simulations, showing that the influence of the electrostatic interaction is considerably larger than that of the VdW interaction. II. SIMULATION DETAILS Classical MD simulations were performed with the NAMD432.10 simulation package. An amorphous cylindrical nanopore of roughly 4 nm of diameter and silica slabs of different thicknesses were fabricated in our group. To create the pore, a crystalline cell of SiO 2with a box length of approximately 6 nm was built, the system was melted at 5000 K and cooled to room temperature with the method described in Ref. 44, and then a cylindrical cavity of ∼4 nm diameter was cut. The process of fabrication of the silica cylindrical pore and the silica slab is explained in detail in a separate paper.44The surface concentration of hydroxyl groups on the surface is 7 .5 nm−2corresponding to highly hydrated silica surfaces.34The silica slabs were created following a similar procedure.TABLE I. LJ potential parameters for silica interaction centers. Parameters Sites σ(nm) ε(kcal/mol) q (e) BZ34Si 0.0178a0.000 00 1.283 OSi 0.27 0.457 056 94 −0.629 OH 0.3 0.457 056 94 −0.533 H 0.0178a0.000 00 0.206 GT36Si 0.25 0.000 1 1.28 OSi 0.27 0.457 −0.64 OH 0.307 0.17 −0.74 H 0.1295 0.000 365 7 0.42 aValues were not mentioned in the model and were chosen arbitrarily small. The silica nanopore and slab contain two types of oxygen atoms depending on the number of silicon atoms to which they are connected. There are bridging oxygens (O Si) bonded to two adjacent silicons and nonbridging oxygens (O H) on the surface attached to only one silicon. Hydrogen atoms are attached to the O Hin order to form the silanols groups (SiOH, Si(OH) 2). The bonded interaction parameters for silica atoms were obtained from Hill and Sauer.45Apart from the hydrogen atoms of the silanols groups that are allowed to rotate, all atoms in the silica pore and slab are immobile, constrained to a fixed position, whereas water molecules are free to move within the pore. Liquid water is defined using 2 models: a set of 3 rigid sites given by the SPC /E46model and a set of 4 sites provided by the TIP4P200547model. The atoms of the silica substrate are allowed to interact with the water sites by means of the Coulomb potential and LJ potential in Eq. (1), ULJ=4ϵ* ,σ12 r12 i,j−σ6 r6 i,j+ -(1) which implements the VdW interaction. LJ parameters and fractional charges for the SiO 2sites are given in Table I. All simulations were made with the NVT ensemble with a fixed room temperature T=298 K using a Langevin thermostat43with a coupling coe fficient of 1.0 ps−1and with the hydrogen atoms included in the Langevin dynamics. An integration time step of 2 fs was utilized and the simulations were run for at least 20 ns. Periodic boundary conditions were set for the simulations of water in the nanopore allowing the calculation of the long-range Coulombic electrostatic interactions with the particle-mesh Ewald sum, using a cut-o ff of 1.2 nm. No periodic boundary conditions were defined for the simulation of water wetting silica slabs in order to allow the calculation of the full electrostatic and VdW interactions between all the water droplet atoms and silica slab atoms. III. RESULTS I: WATER IN PARTIALLY FILLED SILICA PORES In the following, we investigated the configuration of water in a partially filled silica pore for the two di fferent models using MD simulations. Furthermore, we performed an analytical calculation of the di fferent possible phases of water in a cylindrical pore that allows us to interpret the findings.
10.1063_1.4962516
page_0002
114901-3 Pafong, Geske, and Drossel J. Chem. Phys. 145, 114901 (2016) FIG. 1. Top view of the two initial con- figurations used in the MD simulations, labelled in the subsequent figures with “center” (a) and “surface” (b). Here the hydration level is 30%. Si, O Si,H, H are drawn in yellow, blue, and green, while water atoms O, H are indicated by red and white, respectively. The pic- tures were generated using the VMD program.48 A. MD simulation results We evaluated the equilibrated configurations of water in the pore for the GT and BZ surfaces, using di fferent filling ratios and di fferent starting configurations. The pore filling ratios are in the range 30%–97%, based on the estimated number of molecules for 100% filling ratio, which is 2700.30The equilibrium configurations were analyzed by calculating the radial density profile, the distribution of water molecules on the interior pore surface, and the number of hydrogen bonds among water molecules and between water and silica molecules. In order to see how far the final configurations depend on the initial configuration, we used the two di fferent initial configurations shown in Fig. 1, where water is concentrated around the cylinder axis and at the pore surface, respectively. There is thus a void between the water droplet and the silica surface in the first configuration, and a void in the pore center for the second configuration. Fig. 2 shows the radial density profile of water molecules inside the pore as a function of the distance to the pore center, averaged over 15 ns after at least 5 ns of equilibration for each simulation. One can see that water is closer to the GT surface. The GT density profile shows only one peak for a filling ratio of 30%, indicating that all water molecules are in contact with the pore surface. Only after the first layer is completed, a second layer is formed, as is visible for the curves for filling ratios between 40% and 55%. At 65% filling ratio, thewater molecules can also be found in the interior of the pore, indicating a configuration with a completely wetted surface and a compact water droplet in the pore interior. The density profiles for the BZ surface show several layers of water, with a peak height that depends on the filling ratio. Furthermore, the density profile depends on the initial configuration for intermediate filling ratios, with the initial configurations at the surface leading to final configurations with rather flat density profiles. This suggests that for the initial configuration at the boundary, the water droplet forms a “plug” in the pore interior, while for the initial configuration in the center, water forms some type of droplet sitting at the surface. Since the pore surface is rough, some water molecules can also be found inside the silica pore material. In order to test the intuition obtained for the water configurations based on the density profiles, we evaluated the distribution of water molecules within a distance of 0.3 nm of the surface. Fig. 3 shows the resulting surface density profiles, using cylinder coordinates. This figure confirms that for the GT surface, the water droplet first wets the surface completely, before filling the interior. For the BZ surface, the pore surface is only partially covered with water, and the final configuration depends on the starting configuration for intermediate filling ratios. For instance, for 40% filling ratio with the “center” starting configuration, water molecules are concentrated in one angular segment of the surface but are covering the whole length, while for the “surface” starting configuration they FIG. 2. Radial density profile of water in the silica pore for di fferent filling ratios and starting configurations, for the (a) BZ surface and the (b) GT surface. The radius of the pore is 2 nm. The gray area indicates the silica pore surface and is arbitrarily scaled for a clear visibility.
10.1063_1.4962516
page_0003
114901-4 Pafong, Geske, and Drossel J. Chem. Phys. 145, 114901 (2016) FIG. 3. The water density at the pore surface, showing water molecules that are within a distance of 0.3 nm of the silica surface. For the GT surface ((e)-(h)), only filling ratios of 30% and 40% are shown, since for larger filling ratios the surface is completely wetted. For the BZ surface, the top ((a)-(b)) graphs correspond to the “surface” starting configuration and the bottom graphs ((c)-(d)) correspond to the “center” starting configuration, (i) represents the density at 65% filling ratio. occupy only part of the zrange, but all angles. These final configurations for intermediate filling ratios are in fact very plausible if one tries to imagine how the initial configurations can evolve with time in a situation where the water-surface interaction is not strong enough that the entire surface is wetted. When the initial configuration has a water cylinder in the pore center, the entire cylinder gets attracted by the silica molecules under the influence of electrostatic and VdW interaction and moves as a whole towards the pore surface, wetting a specific angular region of the surface. When the initial configuration sits at the pore surface, the water film may rupture along an angular line, and the water will contract to form a plug. Even if one of the two final configurations has a lower free energy, this free energy di fference will not be large, and the transition between them will involve a barrier that is so large that it is not overcome during the simulation time. When the filling ratio is lower (as can be seen for 30%), the plug is not observed for either initial configuration, indicating that there is only one stable configuration. The two di fferent final configurations merge also for larger filling ratios (as can be seen for 65%), where the void left by the water droplet takes the shape of a droplet that sits at the pore surface. Finally, we evaluated the average number of hydrogen bonds formed between water molecules, and between water molecules and silica molecules. This shows to what extent the stronger hydroa ffinity of the GT model a ffects the formationof molecular bonds. We considered two oxygen atoms to be connected via a hydrogen bond if the angle between the intramolecular O—H vector and the intermolecular O ···O vector is less than 30◦, provided that the O ···O separation is less than 0 .335 nm. The results are shown in Fig. 4. For the BZ surface, the number of hydrogen bonds between water molecules reaches the bulk value in the inner part of the pore for a filling ratio larger than 40% with the “surface” initial configuration. Also for the “center” initial condition, the bulk value is reached for the water molecules in the interior of the water droplet. In the GT surface, the bulk value is reached only for filling ratios above 60%. This illustrates the fact that the BZ surface disrupts the water structure more than the GT surface. Accordingly, the number of hydrogen bonds formed between the silica surface and the water molecules is larger for the GT surface. For both models, the maximum number of water-silica hydrogen bonds is already reached at 40% filling ratio, confirming that one and half layer of water molecules is sufficient to completely wet the GT silica surface. It is at first surprising that for the BZ surface, the number of water-silica hydrogen bonds does not increase for filling ratios larger than 40% and stays considerably below the value of the GT surface. This can only be explained by di fferent water orientations near the surface in the two models. In the supplementary material we show that near the BZ surface the OH bonds of
10.1063_1.4962516
page_0004
114901-5 Pafong, Geske, and Drossel J. Chem. Phys. 145, 114901 (2016) FIG. 4. Average number of hydrogen bonds per water molecule for water–silica ((a)-(b)) and water–water ((c)-(d)) contacts, for di fferent initial conditions and filling ratios, as indicated in the legends. The results for the GT silica surface are given on the left-hand side, and those for the BZ silica surface on the right-hand side. water molecules have a preferred orientation, while this is not the case near the GT surface. This means that only part of the water molecules can act as a hydrogen bond donor or acceptor near a BZ silica molecule, while near the GT surface all the water molecules can share hydrogen bonds with the surface atoms. B. Theoretical evaluation of the phase diagram In order to better understand the dependence of the water droplet configuration in the pore on the interaction energies and the filling ratio, we performed a theoretical analysis that is based on surface energy minimization. Denoting the surface area between the water droplet and vacuum with A1, the surface area between the water droplet and the pore material with A2, andγ1as the surface tension between water and vacuum,γ2as the di fference between the surface tension of silica and water with the surface tension of silica and vacuum. The total surface energy of the wetting droplet can be written as ES=γ1·A1+γ2·A2. (2) If we assume that the entropy does not change much between different phases, the configuration of the water droplet in the pore can be obtained by minimizing ESfor a given filling ratio. In order to perform the calculation mostly analytically, we approximated the di fferent possible phases using simple geometrical shapes, so that the energy minimization can be performed by varying at most 2 parameters that characterize the phase. We fixed the ratio of the pore radius and pore length to the value R/L=2/6.1 used in the simulations. We determined the phase diagram in dependence of the filling ratio and the ratio between the two surface energies. We allowed for hydrophilic ( γ2<0) as well as for hydrophobicsurfaces (γ2>0). The surface tension γ1is a positive quantity. Fig. 5 shows the eight phases and the phase diagram obtained from minimizing ES. Phases 1, 2, and 4 are translationally invariant along the cylinder axis and represent the cases of partial, full, and no wetting of the silica surface. Phase 3 represents a plug in the shape of a cylinder that is shorter than the pore. Phases 5-8 describe the cases where the water or the vacuum forms a spherical droplet in the interior or one that intersects with the pore surface. The calculation of the surface energy for phases 1-6 is a straightforward analytical calculation. For phase 1, we had to use mathematica to evaluate the final expression. In order to evaluate phases 7 and 8, we had to resort to a numerical evaluation. We first created a database by FIG. 5. (a) The eight di fferent phases used for energy minimization and (b) the phase diagram in dependence of the ratio of the water-vacuum and water-silica surface energies and of the filling ratio. The aspect ratio r/l between radius and length of the cylindrical pore is 2 /6.1.
10.1063_1.4962516
page_0005
114901-6 Pafong, Geske, and Drossel J. Chem. Phys. 145, 114901 (2016) calculating numerically the volume and surface of the cut droplet for over 1 ×106different combinations of the sphere radius and the distance of the sphere center from the cylinder axis. Then, we used this database to find for a given filling ratio the values that minimize ES. The phase diagram shows clearly three qualitatively different regions that depend on the ratio γ2/γ1. For γ2/γ1.−1, the energy is minimized by having maximum surface area with the pore surface. The water wets the pore completely. Correspondingly, the phases 2 and 6 occur. With increasing filling ratio, the volume of the vacuum becomes smaller, and eventually a free volume that does not touch the surfaces fits into the pore. For su fficiently large filling ratio, the droplet clearly has the smaller surface area with the vacuum and therefore the lower energy. (Our simple calculation did not take into account that the droplet could be stretched, and therefore the position of the phase boundary between phases 2 and 6 should in fact be at a lower filling ratio.) Forγ2/γ1&1, the pore material is highly hydrophobic, and the phases 4 and 5, which have no contact between water and silica, have the lowest energy, depending on the filling ratio. The phase boundary between phases 5 and 4 moves downwards with decreasing aspect ratio r/l, because a vacuum droplet that does not touch the pore wall takes a smaller proportion of the total volume when r/lis smaller. (See the supplementary material for phase diagrams with other aspect ratios.) In the intermediate parameter region −1.γ2/γ1.1, we observe phases 7, 1, and 8 as the filling ratio is increased. These are the phases that have surfaces with the pore and with the vacuum. Since the (absolute value of) water-vacuum energy is larger than that of the water-silica energy, these phases are to a large extent a ffected by the condition that the water-vacuum interface shall be minimum. The transition from phase 7 to phase 1 occurs for lower values of γ2/γ1at smaller filling ratios than for larger γ2/γ1, because a larger surface area to the pore is energetically favorable for negative γ2. For the same reason, phase 8 wins over phase 1 for high filling ratios and negative γ2, because phase 8 has more surface area between water and the pore. The phase boundary to phase 8 moves upwards and the boundary to phase 7 moves downwards with decreasing aspect ratio r/l, because droplets that touch the pore wall only at one side take a smaller proportion of the total volume when r/lis smaller. The droplet phases will also vanish when the ratio r/lbecomes large, as the system then is e ffectively two-dimensional and shows only the three phases that are translationally invariant along the z-axis. (See supplementary material for phase diagrams with other aspect ratios.) Phase 3 does not occur in the phase diagram. It will certainly occur when the aspect ratio r/lbetween the radius and length of the pore becomes smaller, because it has then smaller surface area than phase 1. (See the supplementary material.) In our simulations with the BZ potential, we found this phase for intermediate filling ratios, where it coexists with phase 1. Phase 3 thus might well be metastable. On the other hand, it is also possible that phase 3 is indeed stable in part of the phase diagram due to entropic e ffects, which were not taken into account when calculating the phase diagram. Since in the canonical NVT ensemble the free energy F=E−T Shas to be minimized, phases with larger entropy become more favored when entropy is taken into account. This will shift the phase boundaries somewhat. For instance, when phase 2 contains only two layers of water molecules, its entropy per molecule is smaller than in bulk water. Similarly, the entropy per water molecule is larger in phase 3 than in phase 1, since the water in phase 3 is more bulk-like. With the insights gained from these analytical calcula- tions, we can interpret the results of the MD simulations shown in Figs. 2–4: For the BZ surface, the water wetted the silica surface only partially for all simulated filling ratios, and we observed the phases 1, 3, and 8 depending on the filling ratio. The transition to phase 8 occurs at a filling ratio of approximately 60%. For smaller filling ratios below 15%, we also see phase 7 (see the supplementary material). This means that the ratio of surface energies γ2/γ1is in the interval (−1,0). (Since the surface is hydrophilic, we have γ2<0.) With the GT surface, we observed a complete wetting of the silica surface (phases 2 and 6) for all simulated filling ratios, with a transition between these two phases at a filling ratio around 60%. This means that γ2/γ1<−1. This appears to be the more realistic scenario, as it agrees well with experimental results.2 In order to obtain an additional perspective on the di fferent interaction between water and a silica surface in the two models, we will in Sec. IV investigate the contact angle of water on top of a flat silica slab using both models. IV. RESULTS II: WATER ON TOP OF A SILICA SLAB A good tool to examine the performance of silica potentials is the evaluation of the contact angle of a water droplet wetting the surface. We performed MD simulations of a water droplet on a flat surface of amorphous silica and measured the contact angle. We did not use periodic boundary conditions in order to remove the influence of the neighbouring water periodic images. Instead, the Coulombic and the VdW (LJ) interaction energies between all atoms in the water droplet and the silica slab were calculated exactly. In order to evaluate the contact angle θ, the density profile of all horizontal water layers of 0 .05 nm thickness was determined, and from these a contour plot of the density was obtained. The contour plot was fitted to a circular segment, and the contact angle was deduced from the tangential line to the base of the circular segment. The result is shown in Fig. 6 for both types of potentials, with a slab of thickness t=2.5 nm. For the GT model, the droplet covers the entire surface and has a very small contact angle of 7◦. When we performed the same simulation with periodic boundary conditions, the water layer became completely flat. In contrast, the contact angle of the water droplet on top of the BZ surface is 25◦. These results confirm the findings of Subsection III B, that the GT silica surface is so hydrophilic that water wets it completely, while the BZ surface is less hydrophilic. The contact angle is closely related to the surface tensions that we used for evaluating the phase diagram. The condition that the total surface energy ( ES) of water wetting a silica surface must be minimal for an equilibrated droplet of constant
10.1063_1.4962516
page_0006
A&A 641, A54 (2020) Fig. 6. Abundances relative to H2of MF, DE, F, and EC (Table 10) as a function of the total luminosity/mass ratio (see Sect. 6.3 for details) for the sources of our sample. The linear best-fit to the data is shown for each molecule. luminosity and is not affected by any distance-induced observa- tional bias, since we have checked that molecular abundances are independent from source mass and distance. PC and most HMS- FRs are in very good agreement with the trend observed in our sample, while hot corinos, IMSFRs and some of the other HMS- FRs show slightly higher values. Since the results of the latter are based on interferometric data, this discrepancy could be due to the different angular resolution. Although we accounted for beam dilution effects as consistently as possible (see Sect. 4), lower resolution (single-dish) observations may still result in slightly underestimated molecular column densities. Figure 8 summarises the main result of this analysis, showing the average abundances of the four molecules with respect to the evolutionary stage of the sources. For molecules detected at mul- tiple stages (MF, DE, and EC), average values increase with the evolution, namely from protostellar to intermediate until UCHII regions, preserving the mutual molecular ratios. The increasing trend is particularly evident for MF and DE. Average abundances increasing with time were also found by Gerner et al. (2014) for less complex molecules CH 3OH(methanol), CH 3CN(methyl cyanide), and other simpler molecules, and were predicted by Choudhury et al. (2015) for COMs including MF and DE through evolutionary models of HMCs. 6.4. Implications for the chemistry of COMs The abundances of MF, DE, and EC are very well correlated (r0:92, Fig. 2) and their mutual molecular ratios are nearly constant (Figs. 3–4). The result is very robust since it is based on a sample with good statistics (20 sources in our sample plus 59 sources from literature overall), covering several orders of magnitude in abundance and source luminosity. In some cases, this may indicate a chemical link between the species. This is most likely the case of MF and DE, show- ing the strongest correlations in many parameters (abundance, source size, and FWHM) and a constant 1ratio over a remark- able9orders of magnitude in source luminosity (Fig. 3, upper panel), with a limited scatter both in a large sample of low- to high-mass star-forming regions and among different interstellar environments (Fig. 4). The link may consist in a common forma- tion pathway or in one species being the precursor of the other. Fig. 7. Same as Fig. 6, but for individual molecules MF ( upper panel ), DE (middle panel ), and EC ( lower panel ). The evolutionary classifica- tion is shown for the sources of our sample (different colours), while black symbols represent different interstellar sources taken from litera- ture for comparison (see Table F.1 for references). The black lines fit the data of the sources included in this work. The first scenario is indeed predicted by the theoretical model of Garrod & Herbst (2006) and Garrod et al. (2008), who propose a common formation route through surface chemistry on dust grains at low temperatures ( 50K), from the methoxy precursor CH 3O(see also Allen & Robinson 1977): CH 3O+HCO!CH 3OCHO; (MF) CH 3O+CH 3!CH 3OCH 3: (DE) A54, page 12 of 25
10.1051_0004-6361_202038212
page_0011
A. Coletta et al.: Evolution of COMs in star-forming regions Fig. 8. Average abundances relative to H2(with respective standard errors) of MF, DE, F, and EC (different colours), as a function of the evolutionary stage. Balucani et al. (2015) present instead a gas-phase route able to efficiently form MF from DE at low temperatures ( 10K) through reactions involving the radical CH 3OCH 2: CH 3OCH 3+F!CH 3OCH 2+HF; CH 3OCH 3+Cl!CH 3OCH 2+HCl; CH 3OCH 2+O!CH 3OCHO +H: In addition, the correlated FWHM of the lines (middle panel of Fig. 5), the similar overall range of excitation temperatures (Sect. 5.3), and the spatial coexistence derived from interfero- metric observations (e.g. Brouillet et al. 2013; Bøgelund et al. 2019; El-Abd et al. 2019) suggest that MF and DE could trace the same gas in various environments and evolutionary stages. However, also in the case of species for which a chemical link is not so clear (EC and MF, or EC and DE, showing slightly higher dispertion in molecular ratios, Fig. 3, bottom two panels) a clear abundance trend is observed. A potential link between these molecules may involve the methyl radical CH 3as a com- mon precursor. EC could indeed form through a sequence of gas-phase and grain-surface reactions mainly involving the CN andCH 3radicals (Garrod et al. 2017). We cannot exclude either the existence of a chemical link with formamide, consistent with the abundance correlations ( >0:9) found in Sect. 6.2.1, but the poor statistics obtained for this molecule prevents conclusive considerations, and needs to be improved by further targeted observations. Although the formation paths of formamide are still under debate (see e.g. Bisschop et al. 2007; Barone et al. 2015; Codella et al. 2017; Skouteris et al. 2017; Ligterink et al. 2018; Quénard et al. 2018; López-Sepulcre et al. 2019), recent works propose that it would form more efficiently on icy dust grains during the cold phases of star formation (Jones et al. 2011; López-Sepulcre et al. 2015; Fedoseev et al. 2016). It has to be noted, however, that abundance correlations between molecules do not necessarily imply the existence of a chemical link, as recently proved by Quénard et al. (2018) for formamide andHNCO (isocyanic acid), and confirmed by Belloche et al. (2020) in a sample of hot corinos. These observational corre- lations seem to be a necessary but not sufficient condition to claim a chemical link. Nevertheless, observations are needed to test models and understand how molecules are formed. This work shows, in fact, that between molecules whose chemistry isexpected to be related (such as MF and DE) the correlations are tighter. Furthermore, a clear trend of increasing molecular abun- dances with L=M(mainly governed by L) emerges for all species, spanning up to4orders of magnitude in abundance and 6in L=M, which implies also a trend with the evolutionary stage of the sources (Figs. 6–8). Besides suggesting potential individual links between the COMs, these results allow us to formulate a general, most likely scenario for their formation and evolution. The fact that the molecular ratios are nearly constant across the whole star forma- tion process and among different types of sources is particularly interesting, because the physical conditions in these environ- ments (especially in the case of MF/DE, Fig. 4) are different: pre-stellar cores, shock-dominated regions (protostellar shock and GC clouds), thermal-dominated regions (cores in low- to high-mass star-forming regions), and comets (whose chemical composition is thought to be presolar, see e.g. Rivilla et al. 2020). This seems to reveal a rather universal chemistry for COMs, mainly developed at the cold earliest stages of star for- mation and then essentially preserved through the evolution, being only marginally altered by the evolving physical condi- tions. In more detail, molecules may be formed in pre-stellar cores, possibly in gas phase or on the surface of dust grains, from which they can desorb thanks to non-thermal mechanisms such as cosmic rays (see e.g. Shingledecker et al. 2018; Bonfand et al. 2019; Willis et al. 2020). This would explain the detec- tion and the relative (low) abundances in the pre-stellar cores and the comets. The lack of molecular detections (at least at 2mm) among our 11 HMSCs may be due to the fact that they are tipically much more distant than the observed PCs (which can be resolved even with relatively low resolutions, see e.g. Jiménez-Serra et al. 2016), and thus more affected by beam dilution. Later on, in star-forming regions and GC molecular clouds, other mechanisms are able to massively (and more effi- ciently) desorb the molecules formed on grains: thermal heating and shock-induced heating. This has the effect to significantly increase the observed gas-phase molecular abundances and thus the expected number of detections. This scenario is consistent with the trend we find between abundances and L=M(proxy for the evolutionary stage), as well as with the number of detections we report for each evolutionary group (Sect. 5.1). Moreover, while low luminosity sources (pre-stellar and hot corinos) are usually isolated (or at most binary) systems, high-mass star- forming regions are clustered environments. In these regions, the thermal and shock energy injected to the medium strongly increases with time due to the protostellar activity (heating and protostellar outflows). This produces more and more desorption, accordingly increasing the gas-phase abundances of COMs with evolution. Therefore, the proposed scenario supports the forma- tion of COMs on grain surfaces, indicating that the majority of COMs observed in star-forming regions could be produced by the desorption from icy grain mantles. However, it is still possi- ble that gas-phase formation pathways (see e.g. Balucani et al. 2015; Codella et al. 2017; Skouteris et al. 2019), though not expected to significantly affect the molecular ratios (based on our results), could contribute to the abundance of COMs in cold regions. Moreover, our results suggest that O- and N-bearing COMs may behave similarly in star-forming regions at all stages, shar- ing the same physical conditions (or even direct chemical links) for their formation. This has been found also by Fontani et al. (2007) in hot cores, whereas other authors noticed differences between O- and N-bearing COMs in both the spatial distribution (e.g. Liu 2005; Csengeri et al. 2019) and the radial velocities A54, page 13 of 25
10.1051_0004-6361_202038212
page_0012
A&A 641, A54 (2020) (Blake et al. 1987). We also note that, given the increasing abun- dance trend, molecular destruction routes seem to be less effi- cient than formation/desorption mechanisms, especially at later evolutionary stages (i.e. higher luminosities). However, destruc- tion routes represent a less investigated but non-negligible topic, as they can in principle affect the predicted molecular abun- dances (see e.g. Garrod 2013; Shingledecker et al. 2019; Ascenzi et al. 2019 and refs. therein). Lastly, we stress that the angular resolution of our data (Table 2) is larger than the size of the observed sources. Although this issue has been taken into account through the beam dilution factor applied in the line fitting procedure (see Sect. 4), we are still not able to spatially resolve the inner struc- ture of the targets, which is often fragmented into multiple smaller objects in potentially diverse evolutionary stages. The observed emission could hence include contributions from both the inner hot core and its cooler outer envelope, preventing a clear distinction between nearby emission zones, and causing sometimes potentially misleading correlations among differently distributed molecules. High angular resolution interferometric observations would be able to confirm more robustly the pro- posed scenario for the formation of COMs, as they can more accurately identify spatial correlations and resolve the poten- tial protostellar multiplicity within a region (see e.g. Murillo et al. 2018). Nevertheless, we do not find relevant differences by comparing our results to interferometric data from literature, seemingly indicating that the observed chemistry is almost the same across different spatial scales within star-forming regions. 7. Summary and conclusions In this work we have analysed spectra at 3,2, and 0:9mm of 39 selected high-mass star-forming regions at different evolution- ary stages (HMSCs to UCHIIs) obtained with the IRAM-30m telescope, searching for rotational transitions of the complex O- bearing molecules CH 3OCHO (MF) and CH 3OCH 3(DE), and N-bearing molecules NH 2CHO (F) and C2H5CN(EC). We have reported molecular detections in 20 sources, performing a line fitting procedure to derive the main physical parameters for each molecule. We summarise below the main results of this study: – The highest number of detections was reported in UCHII regions ( 45%, 9 out of 20 sources). DE was detected in 19 sources, while MF in 13, EC in 9, and F in 5. – We observe relevant discrepancies between the total molec- ular column densities obtained at different wavelengths (up to 2 orders of magnitude between 0:9and 3mm), so that in all sources N3(3mm)>N2(2mm)>N1(0:9mm)and N2=N1>N3=N2. This can be interpreted as an effect of the differential attenuation caused by dust opacity at each fre- quency (d/ ), proving that dust properties have indeed to be considered when dealing with young, tipically dust-rich star-forming regions at multiple wavelengths. Therefore, we chose the 2mm data for our analysis (being the band that reported the most detections) and found source-averaged col- umn densities ranging from 1015to1018cm2for MF, DE, and EC, and from 1014to1017cm2for F. – The derived abundances with respect to H2are1010107 for MF and DE, 10121010for F, and1011109 for EC. For all species we find a consistent overall range of linewidths (210km s1) and excitation temperatures (20220K). – We find very strong correlations ( r0:92) between the abundances of MF, DE, and EC, spanning 3orders of magnitude in abundance, uniformly covered by our sample.We have compared our results with heterogeneous sources from literature (including low-, intermediate- and high-mass star-forming regions, a protostellar shock region, pre-stellar cores and Galactic centre clouds), which confirmed and expanded the correlations to 4orders of magnitude in abun- dance for all tracers. We also find nearly constant molecular ratios with respect to source luminosity across all evolution- ary stages and among different types of sources, indicating that the chemistry of COMs is mainly developed at early stages and then preserved during the evolution, barely altered by the changing local physical conditions. These results may suggest a potential link between MF, DE, and EC, whereas for F, though consistent with correlations ( r>0:9), we can- not draw conclusions due to the poor statistics. In particular, we claim that MF and DE are most likely chemically linked, since they show the strongest correlation in most parame- ters (abundance, FWHM, and source size) and a remarkably constant ratio of1across a wide variety of sources at all evolutionary stages (also including comets), spanning a strik- ing9orders of magnitude in luminosity. The link may consist in a common formation pathway, such as from pre- cursor CH 3Oas predicted by Garrod & Herbst (2006) and Garrod et al. (2008), or in one species being the precursor of the other, as proposed by Balucani et al. (2015) with MF forming from DE. MF-EC and DE-EC may share CH 3as common precursor instead (see e.g. Beuther et al. 2007). Although observational correlations alone are not enough to prove a chemical link, this work shows that they are tighter among molecules whose chemistry is expected to be related (e.g. MF and DE). – We have also evaluated the variation of molecular abun- dances with the evolutionary stage of the source (traced by the luminosity/mass ratio) finding a clear increasing trend for all species over up to 6 orders of magnitude in L=M, ranging from pre-stellar cores and hot corinos to UCHIIs. – Based on correlations, molecular ratios and evolutionary trend, we propose a general scenario for the formation and evolution of COMs, which involves a prevalent formation at low temperatures in the earliest phases of star forma- tion (likely mainly on frozen dust grains) followed by a growing desorption powered by the progressive thermal and shock-induced heating of the core with evolution. This would explain the increasing observed gas-phase abundances and number of molecular detections. Moreover, these results sug- gest that O- and N-bearing COMs might have a similar behaviour in star-forming regions at all stages. Interestingly, this analysis also points out that molecular abundances might serve as evolutionary tracers within the whole star formation process. In conclusion, we stress that the physical parameters derived in our sample represent average values across the whole clumps, and could therefore include also contributions from outside the cores. Relevant improvements to this work will come from high angular resolution observations, able to resolve the inner struc- ture of these regions and hence to better locate the molecular emission, allowing to more accurately identify spatial correla- tions between COMs. In particular, interferometric observations of a large sample of star-forming regions in different evolution- ary stages, like the one studied in this work, will be able to confirm and improve the proposed scenario for the formation and evolution of COMs. Acknowledgements. We thank the IRAM-30m staff for the precious help during the different observing runs. V.M.R. has received funding from the European A54, page 14 of 25
10.1051_0004-6361_202038212
page_0013

Dataset Card for Dataset Name

Dataset Summary

This dataset card aims to be a base template for new datasets. It has been generated using this raw template.

Supported Tasks and Leaderboards

[More Information Needed]

Languages

[More Information Needed]

Dataset Structure

Data Instances

[More Information Needed]

Data Fields

[More Information Needed]

Data Splits

[More Information Needed]

Dataset Creation

Curation Rationale

[More Information Needed]

Source Data

Initial Data Collection and Normalization

[More Information Needed]

Who are the source language producers?

[More Information Needed]

Annotations

Annotation process

[More Information Needed]

Who are the annotators?

[More Information Needed]

Personal and Sensitive Information

[More Information Needed]

Considerations for Using the Data

Social Impact of Dataset

[More Information Needed]

Discussion of Biases

[More Information Needed]

Other Known Limitations

[More Information Needed]

Additional Information

Dataset Curators

[More Information Needed]

Licensing Information

[More Information Needed]

Citation Information

[More Information Needed]

Contributions

[More Information Needed]

Downloads last month
0
Edit dataset card