Articles

ACCURATE SPECTROSCOPIC CHARACTERIZATION OF ETHYL MERCAPTAN AND DIMETHYL SULFIDE ISOTOPOLOGUES: A ROUTE TOWARD THEIR ASTROPHYSICAL DETECTION

, , , , , and

Published 2014 November 4 © 2014. The American Astronomical Society. All rights reserved.
, , Citation C. Puzzarini et al 2014 ApJ 796 50 DOI 10.1088/0004-637X/796/1/50

0004-637X/796/1/50

ABSTRACT

Using state-of-the-art computational methodologies, we predict a set of reliable rotational and torsional parameters for ethyl mercaptan and dimethyl sulfide monosubstituted isotopologues. This includes rotational, quartic, and sextic centrifugal-distortion constants, torsional levels, and torsional splittings. The accuracy of the present data was assessed from a comparison to the available experimental data. Generally, our computed parameters should help in the characterization and the identification of these organo-sulfur molecules in laboratory settings and in the interstellar medium.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

Given the improved capabilities of new astronomical observatories in terms of spectral coverage and spatial resolution, the corresponding surveys cover large frequency ranges and can contain many unidentified lines. The latter features usually correspond to "new" molecules and/or rare isotopologues of previously detected species. The "cleaning" of the corresponding spectral confusion implies the full assignment of astronomical spectra, which in turn requires a complete spectroscopic characterization of all of the involved species at different temperatures.

Organic molecules can be detected in low-energy, excited vibrational states in hot molecular cores through their rotational spectra. Furthermore, nonrigid species such as ethyl mercaptan and dimethyl sulfide present internal rotation motions restricted by energy barriers. These motions can produce splitting of the ground and low-excited vibrational levels through the tunneling effect. Their complete characterization at low temperatures requires analysis of the vibrational spectrum in the far-infrared region. Unfortunately, a large number of detectable molecules are not well characterized. For many species, although accurate spectroscopic data are available, they are limited to the vibrational ground state and to the most abundant isotopologues. The effect of temperature and the fact that isotopic composition can be very different from one extraterrestrial source to another are often not considered. However, species containing less abundant isotopes or molecules with populated low vibrational states can play important chemical roles in many sources. An example is provided by methyl formate (a nonrigid molecule), for which investigations have been reported for the rotational spectrum in the vibrational ground state for the main isotopic species (Brown et al. 1975; Churchwell & Winnewisser 1975) as well as for several isotopologues (Carvajal et al. 2009; Margulès et al. 2010; Tercero et al. 2012). Features due to the first excited torsional state have also been detected for the main isotopologue (Kobayashi et al. 2007; Demyk et al. 2008) and 13C-containing (Carvajal et al. 2010) isotopologues. The main isotopic species was also observed in the second excited torsional state (Takano et al. 2012).

State-of-the-art ab initio methods can be used to compute highly accurate spectroscopic constants, even for nonrigid molecules (Brites et al. 2008; Halvick et al. 2011; Puzzarini et al. 2014a; Senent et al. 2009, 2014). Today, these constants are known to provide the required accuracy to guide experimental investigations of rotational and far-infrared spectra. While experimental determinations for rare isotopic species can be hampered by their low natural abundance, computations provide the same accuracy for all isotopologues. Therefore, quantum-chemical calculations are well suited to support and complement experimental and astronomical studies that also involve rare isotopic species. Examples can be found in investigations of the main isotopologues and several other isotopologues containing D and 13C of dimethyl-ether (DME; Senent et al. 2012; Carvajal et al. 2012, 2014) or propane (Villa et al. 2013).

Recently, we conducted a study of the main isotopic species of dimethyl sulfide (DMS) and ethyl mercaptan (ETSH) at low temperatures (Senent et al. 2014) considering their potential astrophysical relevance. In particular, the latter molecule was considered to be a potential detectable sulfur organic compound based on different arguments: the importance of sulfur chemistry in the interstellar medium (Charnley 1997) and because the O-analog of ethyl mercaptan, ethanol, is a well-known astrophysical molecule. We should consider that usually, with few exceptions (Cernicharo et al. 1987), the detection of S-bearing species follows the detection of corresponding O-analogs. The expectations of astrophysicists were recently satisfied when an exhaustive search led to the detection of ethyl-mercaptan in Orion in 2014 (Kolesniková et al. 2014). For many years, methyl mercaptan was the only sulfur nonrigid molecule detected in astrophysical sources (Linke et al. 1979).

Differences in their formation processes have made determination of the isotopic abundances of different sources particularly interesting as a means of investigating interstellar and galactic evolution (Mauersberger et al. 2004). In addition, fluctuations of the relative isotopic abundances are tools for identifying chemical, geophysical, and biological processes (Canfield 2001; Farquhar & Wing 2003; Mauersberger et al. 2004). Hence, relative isotopic abundances have been estimated for meteorites, the moon, cosmic rays, and stars (Mauersberger et al. 2004). There is some evidence demonstrating that isotope yields follow different synthesis procedures. Whereas 32S, 34S, and 33S are the primary products of oxygen burning in a star, 36S is produced when the primary sulfur isotopes capture neutrons during helium and carbon burning (Mauersberger et al. 1996).

Considering their potential importance, in the present work, we investigated several rare monosubstituted isotopologues of ethyl mercaptan and dimethyl sulfide containing either a rare S-isotope or deuterium. Sulfur, the tenth most abundant cosmological element, presents four stable isotopes 32S, 34S, 33S, and 36S whose abundance ratios in the solar system were estimated to be 95.02%, 4.21%, 0.75%, and 0.021% (Anders & Grevesse 1989), respectively. Their relative abundances in Orion KL were determined to be 32S/34S = 20 ± 6, 32S/33S = 75 ± 29 (Tercero et al. 2010a, 2010b). The first molecule containing 36S,C36S, was detected in Galactic molecular hot cores by Mauersberger et al. (1996), who estimated a relative abundance of 34S/36S = 115 ± 17, which is smaller than that in the solar system. This ratio is consistent with what was later determined by Mauersberger et al. (2004), 107 ± 15, in the carbon star IRC+ 10216 based on the detection of the rotational transitions of C36S and Si36S. Moreover, several S-containing molecules were detected in the ISM raging from diatomics (e.g. NS, SO, and SH) to more complex small organic molecules (e.g. C2S, H2S, OCS, C3S, H2CS, HSCN, HNCS, CH3CS, etc.; Ziurys 2006; Millar et al. 1986; Irvine et al. 1988; Goldsmith et al. 1981; Minh et al. 1991; Frerking et al. 1979; Linke et al. 1979). Si36S is the first compound containing 36S detected in a star (Mauersberger et al. 2004).

In the present work, by following the procedures of Senent et al. (2014), we have investigated the 34S-, 36S-, 33S-, and deuterium-containing ETSH and DMS species, considering them not only in the vibrational ground state, but also in their excited torsional states because at the temperature of the hot molecular cores, these excited states and their splittings can be populated. Hence, we provide a detailed spectroscopic characterization of these S-containing molecules and isotopologues to help their detection in the ISM. Thus, the aim of this paper is to predict the isotopic substitution effect on the rotational and torsional parameters. To assess their accuracy, the computed properties were compared with the available experimental data, in particular, for the main isotopologues (Senent et al. 2014). While for the latter we refer to Senent et al. (2014) for an account on previous theoretical and experimental studies, here we mention that for other isotopic species, only a few studies are available (Hayashi et al. 1989; Schmidt & Quade 1975; Wolff & Szydlowski 1985; Manocha et al. 1973; Kretschmer et al. 1995; Kolesniková et al. 2014).

2. METHODOLOGY

In this study, we followed the methodologies described in Puzzarini et al. (2010), Puzzarini (2013), and Senent (1998a, 1998b, 2001). In particular, we refer interested readers to Puzzarini et al. (2010) and Puzzarini (2013) for rotational spectroscopy and to Senent (1998a, 1998b, 2001) for torsional analysis. In particular, concerning the spectroscopic characterization of ETSH and DMS, all computational details can be found in Senent et al. (2014). In the following, only a brief summary is provided. Here, we also point out that DMS has two equivalent methyl groups leading to nine equivalent minima in the potential energy surface (PES), while ETSH has a unique methyl group which is responsible for three equivalent minima. In addition, for the latter, a second torsional coordinate, the thiol torsion (SH torsion), leads to two conformers, the gauche and trans forms. Then, the coupling of the two torsional motions in ETSH (methyl and thiol torsions) generates a PES with nine minima (for further details, see Senent et al. 2014).

2.1. Rotational Spectroscopy

To obtain accurate equilibrium rotational constants, the equilibrium structures of ETSH and DMS were determined by means of a composite scheme (see Senent et al. 2014) which is based on additivity at an energy-gradient level (Heckert et al. 2005, 2006) and employs the coupled-cluster singles and doubles approximation (CCSD) augmented by a perturbative treatment of triple excitations [CCSD(T)] (Raghavachari et al. 1989) in conjunction with correlation-consistent basis sets, cc-p(C)VnZ (n = T, Q, 5) (Dunning 1989; Woon & Dunning 1995). Second, to derive vibrational ground-state and torsional-excited-state rotational constants, the equilibrium rotational constants were corrected for vibrational effects with the corresponding corrections being obtained by means of second-order vibrational perturbation theory (VPT2; Mills 1972) at the MP2/cc-pVTZ, CCSD/cc-pVTZ, and CCSD(T)/cp-VTZ levels (for details, see Senent et al. 2014), where MP2 stands for the Møller–Plesset theory to the second order (Møller & Plesset 1934). These calculations implied the evaluation of cubic force fields, and therefore they also allowed us to determine quartic and sextic centrifugal-distortion constants. Different levels of theory were considered in order to verify the cheapest one while still providing reliable and accurate results. All of these calculations were carried out with the quantum-chemical CFOUR program package (2012).

To further improve the predictive capabilities of our computed parameters (i.e., rotational and centrifugal-distortion constants), an empirical scaling procedure was employed for the ground-state parameters. For a generic parameter X, the latter procedure is based on multiplying the computed value of X for a monosubstituted isotopologue (denoted by the superscript iso) for the corresponding experiment/theory ratio for the main (32S-containing) isotopic species (denoted by the superscript main):

Equation (1)

where scal, exp, and calc denote the scaled, experimental, and quantum-chemically calculated values for X, respectively. This approach is extensively used in the field of rotational spectroscopy, and its validity has been discussed, for example, in Puzzarini et al. (2012). For gauche-ETSH, for which a complete set of experimental rotational and quartic centrifugal-distortion constants is available for the 34S-containing species, the latter was used for the reference isotopologue in order to derive the scaled parameters for the main isotopic species.

2.2. Torsional Analysis

For the analysis of the vibrational spectrum in the far-infrared region, we have performed a torsional analysis following the theoretical methodology described in our previous paper (Senent et al. 2014). The energy levels are determined variationally by solving the following two-dimensional Hamiltonian:

which depends on two independent coordinates qi and qj. For ETSH, qi and qj are identified as the CH3 torsion (θ) and the SH torsion (α), respectively. In DMS, both coordinates correspond to methyl internal rotations (θ1 and θ2). V(qi, qj) represents the two-dimensional potential energy surface (2D-PES). VZPVE(qi, q2), Bqiqj, and V ' (qi, qj) denote the zero-point vibrational energy correction, the kinetic energy parameters, and the Podolsky pseudopotential, respectively. For their definition, the readers are referred to Senent (1998a, 1998b).

Since the 2D-PES is isotopically invariant, here we use the surfaces generated in our previous study to treat the main isotopologue. These surfaces were obtained from the total electronic energies calculated at the CCSD(T)/aug-cc-pVTZlevel of theory (Kendall et al. 1992) for a number NS of selected geometries defined for different values of the independent coordinates (NS = 26 for ETSH and NS = 7 for DMS). For each point of the PES, the remaining 3Na-6-n internal coordinates (Na = number of atoms, n = 2 dihedral angles) were optimized at the CCSD/aug-cc-pVTZ level.

VZPVE(qi, q2), Bqiqj, and V '(qi, qj) are, however, isotopically dependent. For each NS structure and for each isotopologue, Bqiqj and V '(qi, qj) were determined using the code ENEDIM, also employed in the variational calculation of the energy levels. VZPVE(qi, q2) was calculated within the harmonic approximation at the MP2/aug-cc-pVTZ level using the Gaussian 09 package. For all details, the reader is referred to Senent et al. (2014).

3. RESULTS

3.1. Rotational Parameters

The computed vibrational-ground-state rotational and centrifugal-distortion constants of the isotopologues considered for gauche-ETSH, trans-ETSH, and DMS are collected in Tables 13, with the Watson's A reduction in the I r representation (Watson 1977) taken into consideration. Tables 1 and 2 summarize in detail the spectroscopic parameters obtained for the isotopologues considered for gauche-ETSH and trans-ETSH. Table 3 reports the results for DMS isotopic species. While in Tables 13 we only report our best-computed values (i.e., those based on CCSD(T) calculations), the CCSD parameters scaled according to Equation (1), and the available experimental data, the complete set of our results is supplied in our tables. In particular, we present the comparison of the MP2, CCSD, and CCSD(T) results.

Table 1. Computed, Scaled, and Experimental Rotational Parametersa for gauche-ethyl Mercaptan

gauche-CH3CH232SH
Parameter Calculated Scalede Experiment Kolesniková et al. (2014)
  MP2b CCSDc CCSD(T)d CCSD 0+ 0
A0 28783.924 28779.449 28775.409 28748.096 28747.4104(66) 28747.2715(65)
B0 5300.793 5300.891 5298.833 5295.107 5295.1422(36) 5295.0008(36)
C0 4850.955 4850.923 4849.312 4845.829 4845.9421(36) 4845.9689(36)
ΔJ 3.362 3.287 3.364 3.328 3.326369(20) 3.323582(20)
ΔJK −20.330 −19.571 −19.809 −18.402 −18.39280(60) −18.35859(60)
ΔK 205.641 205.263 206.297 206.932 204.1591(82) 203.9217(81)
δJ 0.537 0.518 0.534 0.514 0.514429(12) 0.513170(14)
δK 10.016 9.592 9.960 8.482 8.781(12) 8.555(12)
ΦJ −0.0027 −0.0027 −0.0029   0.0029872(28) 0.0029225(31)
ΦJK −0.1125 −0.1182 −0.1201   0.0790(13) 0.0661(13)
ΦKJ 0.9420 1.0045 0.9973   −1.3341(45) −1.2955(44)
ΦK −3.3603 −3.6061 −3.5997   6.341(48) 5.171(46)
ϕJ −0.0011 −0.0011 −0.0012   0.0012107(15) 0.0011805(16)
ϕJK −0.0597 −0.0631 −0.0647   0.04561(63) 0.04326(65)
ϕK 1.9055 1.9526 2.0270   5.858(95) 4.825(97)
gauche-CH3CH234SH
Parameter Calculated Scalede Experiment Kolesniková et al. (2014)
  MP2b CCSDc CCSD(T)d CCSD 0+ 0
A0 28745.294 28741.143 28737.106 28709.078 28709.699(24) 28708.964(26)
B0 5181.142 5181.156 5179.154 5175.469 5175.5685(36) 5175.4383(36)
C0 4749.762 4749.706 4748.131 4744.843 4744.7050(36) 4744.7310(36)
ΔJ 3.215 3.143 3.217 3.179 3.18351(18) 3.18084(18)
ΔJK −19.798 −19.064 −19.290 −17.899 −17.9512(15) −17.8994(14)
ΔK 204.658 204.289 205.308 203.072 206.8(14) 205.1(14)
δJ 0.502 0.484 0.499 0.480 0.481132(82) 0.480129(71)
δK 9.628 9.214 9.569 8.326 8.271(34) 8.024(35)
ΦJ −0.0024 −0.0024 −0.0025      
ΦJK −0.1066 −0.1111 −0.1129      
ΦKJ 0.8831 0.9400 0.9328      
ΦK −3.1091 −3.4068 −3.4003      
ϕJ −0.0010 −0.0010 −0.0011      
ϕJK −0.0545 −0.0577 −0.0592      
ϕK 1.8209 1.8821 1.9542      
gauche-CH3CH233SH gauche-CH3CH236SH
Parameter Calculated Scaled Calculated Scaled     
  CCSDc CCSD(T)d CCSDe CCSDc CCSD(T)d      CCSDe     
A0 28759.688 28755.646 28727.602 28707.028 28702.999 28675.000
B0 5239.310 5237.282 5233.559 5073.263 5071.311 5067.6946
C0 4798.905 4797.313 4793.991 4658.229 4656.686 4653.4594
ΔJ 3.210 3.288 3.247 3.010 3.087 3.045
ΔJK −19.208 −19.541 −18.034 −18.744 −18.833 −17.599
ΔK 204.239 205.786 203.023 204.271 204.428 203.054
δJ 0.500 0.515 0.496 0.454 0.469 0.451
δK 9.402 9.758 8.496 8.846 9.222 7.994
ΦJ −0.0026     −0.0022    
ΦJK −0.1145     −0.1048    
ΦKJ 0.9711     0.8833    
ΦK −3.5034     −3.2284    
ϕJ −0.0011     −0.0009    
ϕJK −0.0603     −0.0531    
ϕK 1.9160     1.8202    
gauche-CH3CH232SD
Parameters Calculated Scaled Experiment Schmidt & Quade (1975)   
  CCSDc CCSD(T)d CCSDe      
A0 26092.013 26087.580 26062.904 26064.003    
B0 5195.122 5193.214 5189.419 5190.090    
C0 4773.556 4772.141 4768.672 4768.620    
ΔJ 3.378 3.521 3.416      
ΔJK −13.104 −13.931 −12.303      
ΔK 140.806 143.593 139.967      
δJ 0.585 0.626 0.580      
δK 13.190 22.893 11.919      
ΦJ −0.0102          
ΦJK −0.0664          
ΦKJ 0.8960          
ΦK −4.1109          
ϕJ −0.0046          
ϕJK −0.1203          
ϕK 4.7852          

Notes. aRotational constants in MHz; quartic and sextic centrifugal-distortion constants in kHz and Hz, respectively. Watson A-reduction in the I r representation is used. bEquilibrium constants corresponding to the best-estimated equilibrium structure (see text; Senent et al. 2014). Vibrational corrections to rotational constants, and quartic and sextic centrifugal-distortion constants at the MP2/cc-pVTZ level. cEquilibrium constants corresponding to the best-estimated equilibrium structure (see text; Senent et al. 2014). Vibrational corrections to rotational constants, and quartic and sextic centrifugal-distortion constants at the CCSD/cc-pVTZ level. dEquilibrium constants corresponding to the best-estimated equilibrium structure (see text; Senent et al. 2014). Vibrational corrections to rotational constants, and quartic and sextic centrifugal-distortion constants at the CCSD(T)/cc-pVTZ level. eFor 32S, the experimental data of the 34S-containing isotopologue were used for the scaling procedure; for 34S, 33S, 36S, and D, the experimental values of 32S were employed. For experiment, the averaged 0+ and 0 values are used.

Download table as:  ASCIITypeset images: 1 2

Table 2. Computed, Scaled, and Experimental Rotational Parametersa for trans-ethyl Mercaptan

trans-CH3CH232SH
Parameters Calculated Experiment Kolesniková et al. (2014)
MP2b CCSDc CCSD(T)d
A0 28460.983 28460.659 28451.975 28416.7604 (18)
B0 5492.059 5491.347 5489.912 5485.77901 (15)
C0 4886.820 4886.618 4885.345 4881.81770 (15)
ΔJ 3.795 3.679 3.767 3.83217 (27)
ΔJK −23.558 −22.847 −23.107 −22.4549 (58)
ΔK 196.710 197.969 198.873 210.03 (18)
δJ 0.646 0.617 0.634 0.65664 (11)
δK 6.091 5.807 5.967 7.342 (20)
ΦJ −0.0116 −0.0120 −0.0130 −0.00086 (17)
ΦJK −0.0415 −0.0467 −0.0423 0.367 (18)
ΦKJ 0.9175 1.0659 1.0607 −2.717 (106)
ΦK −4.6243 −5.3975 −5.4647 26.7 (80)
ϕJ −0.0057 −0.0059 −0.0064 −0.000935 (91)
ϕJK −0.2014 −0.2136 −0.2265 −0.164 (26)
ϕK −1.0809 −1.2489 −1.3450 28.89 (97)
trans-CH3CH234SH
Parameters Calculated Scaled     
  MP2b CCSDc CCSD(T)d CCSDe     
A0 28374.015 28373.744 28365.113 28329.979
B0 5373.179 5372.489 5371.085 5367.041
C0 4789.981 4789.783 4788.536 4785.077
ΔJ 3.659 3.541 3.634 3.688
ΔJK −23.153 −22.564 −22.585 −22.177
ΔK 195.493 197.344 196.995 209.367
δJ 0.613 0.585 0.602 0.622
δK 5.948 5.624 5.868 7.110
ΦJ −0.0105 −0.0109 −0.0117  
ΦJK −0.0408 −0.0457 −0.0418  
ΦKJ 0.8413 0.9788 0.9735  
ΦK −4.2527 −4.9684 −5.0279  
ϕJ −0.0051 −0.0053 −0.0057  
ϕJK −0.1854 −0.1966 −0.2085  
ϕK −0.9751 −1.1294 −1.2178  
trans-CH3CH233SH
Parameters   Calculated Scaled     
    CCSDc CCSD(T)d CCSDe     
A0   28415.795 28407.139 28371.966
B0   5430.260 5428.841 5424.754
C0   4836.877 4835.617 4832.125
ΔJ   3.610 3.699 3.761
ΔJK   −22.716 −22.838 −22.326
ΔK   197.693 197.905 209.737
δJ   0.601 0.617 0.639
δK   5.716 5.917 7.227
ΦJ   −0.0114 −0.0123  
ΦJK   −0.0462 −0.0431  
ΦKJ   1.0204 1.0182  
ΦK   −5.1738 −5.2391  
ϕJ   −0.0056 −0.0060  
ϕJK   −0.2048 −0.2167  
ϕK   −1.1869 −1.3222  
trans-CH3CH236SH
Parameters   Calculated Scaled     
    CCSDc CCSD(T)d CCSDe     
A0   28296.527 28287.940 28252.881
B0   5265.097 5263.722 5259.759
C0   4702.094 4700.871 4697.475
ΔJ   3.400 3.514 3.542
ΔJK   −22.202 −22.115 −21.821
ΔK   196.451 195.316 208.420
δJ   0.553 0.573 0.588
δK   5.437 5.774 6.874
ΦJ   −0.0099 −0.0106  
ΦJK   −0.0447 −0.0422  
ΦKJ   0.9050 0.9026  
ΦK   −4.6015 −4.6568  
ϕJ   −0.0048 −0.0052  
ϕJK   −0.1820 −0.1926  
ϕK   −1.0252 −1.1520  
trans-CH3CH232SD
Parameters Calculated Scaled Experiment Schmidt & Quade (1975)
    CCSDc CCSD(T)d CCSDe
A0 27202.948 27194.339 27160.989 27155.91
B0 5308.668 5307.475 5303.285 5304.36
C0 4706.623 4705.557 4701.999 4702.60
ΔJ 3.166 3.253 3.298  
ΔJK −16.455 −17.670 −16.172  
ΔK 147.484 154.696 156.469  
δJ 0.524 0.541 0.557  
δK 3.826 3.671 4.838  
ΦJ −0.0417 −0.0452    
ΦJK −0.0860 −0.0814    
ΦKJ 5.2740 5.4727    
ΦK −24.3435 −25.1609    
ϕJ −0.0208 −0.0225    
ϕJK −0.6558 −0.7025    
ϕK −6.3870 −6.8136    

Notes. aRotational constants in MHz; quartic and sextic centrifugal-distortion constants in kHz and Hz, respectively. Watson A-reduction in the I r representation is used. bEquilibrium constants corresponding to the best-estimated equilibrium structure (see text; Senent et al. 2014). Vibrational corrections to rotational constants, and quartic and sextic centrifugal-distortion constants at the MP2/cc-pVTZ level. cEquilibrium constants corresponding to the best-estimated equilibrium structure (see text; Senent et al. 2014). Vibrational corrections to rotational constants, and quartic and sextic centrifugal-ditortion constants at the CCSD/cc-pVTZ level. dEquilibrium constants corresponding to the best-estimated equilibrium structure (see text; Senent et al. 2014). Vibrational corrections to rotational constants, and quartic and sextic centrifugal-ditortion constants at the CCSD(T)/cc-pVTZ level. eEmpirically scaled parameters (see text) starting from those computed as at footnote c.

Download table as:  ASCIITypeset images: 1 2

Table 3. Computed, Scaled, and Experimental Rotational Parametersa for Dimethyl-Sulfide

CH332SCH3
Parameter Calculated Experiment Vacherand et al. (1987) Experiment Hayashi et al. (1989).
  MP2b CCSDc CCSD(T)d
A0 17836.230 17832.456 17825.867 17810.0389(35) 17809.734(8)
B0 7630.756 7631.159 7629.422 7621.12253(110) 7621.098(2)
C0 5725.777 5725.526 5724.127 5717.76282(101) 5717.769(2)
ΔJ 8.692 8.324 8.609 8.04258(118) 8.06(4)
ΔJK −40.406 −38.073 −39.071 −35.2214(103) −35.45(43)
ΔK 140.456 138.442 139.591 139.572(49) 140.88(165)
δJ 3.139 2.970 3.086 2.82209(21) 2.84(2)
δK 3.635 3.396 3.440 3.8458(156) 3.29(49)
ΦJ −0.0457 −0.0485 −0.0519    
ΦJK −0.0436 −0.0571 −0.0528 −0.1152(138)  
ΦKJ 1.3772 1.5930 1.6203 0.462(82)  
ΦK −3.8297 −4.4210 −4.4775    
ϕJ −0.0227 −0.0241 −0.0258    
ϕJK −0.2395 −0.2654 −0.2759 0.2277(161)  
ϕK 0.8114 0.7648 0.7885    
CH334SCH3
Parameter   Calculated Scaled      Experiment Hayashi et al. (1989).
    CCSDc CCSDd CCSDe       
A0   17420.928 17414.496 17399.028 17398.825(9)
B0   7631.361 7629.625 7621.324 7621.335(4)
C0   5682.329 5680.944 5674.624 5674.658(8)
ΔJ   8.274 8.558 8.011 8.46(5)
ΔJK   −37.506 −38.496 −34.697 −35.12(40)
ΔK   134.375 135.486 135.472 17398.825(9)
δJ   2.995 3.112 2.846  
δK   3.198 3.230 3.621  
ΦJ   −0.0472      
ΦJK   −0.0411      
ΦKJ   1.4401      
ΦK   −4.0653      
ϕJ   −0.0234      
ϕJK   −0.2445      
ϕK   0.7414      
CH333SCH3
Parameter   Calculated Scaled      Experiment Kretschmer et al. (1995)
    CCSDc CCSD(T)d CCSDe       
A0   17620.725 17614.217 17598.574 17598.3002(14)
B0   7631.263 7629.527 7621.2264 7622.2643(24)
C0   5703.472 5702.080 5695.7379 5694.7067(24)
ΔJ   8.582 8.582 8.034 8.115(32)
ΔJK   −38.776 −38.776 −34.955 −35.36(16)
ΔK   137.481 137.481 137.462 137.07(27)
δJ   3.099 3.099 2.834 2.845(8)
δK   3.335 3.335 3.728 3.63(25)
ΦJ   −0.0478      
ΦJK   −0.0488      
ΦKJ   1.5135      
ΦK   −4.2362      
ϕJ   −0.0238      
ϕJK   −0.2546      
ϕK   0.7530      
CH336SCH3
Parameter   Calculated Scaled       
    CCSDc CCSD(T)d CCSDe       
A0   17050.624 17044.335 17029.190  
B0   7631.541 7629.806 7621.504  
C0   5642.268 5640.897 5634.618  
ΔJ   8.513 8.513 7.969  
ΔJK   −37.977 −37.977 −34.235  
ΔK   131.786 131.786 131.768  
δJ   3.134 3.134 2.866  
δK   3.024 3.024 3.381  
ΦJ   −0.0460      
ΦJK   −0.0276      
ΦKJ   1.3083      
ΦK   −3.7576      
ϕJ   −0.0229      
ϕJK   −0.2262      
ϕK   0.7187      
a-CH332SCH2f
Parameter   Calculated Scaled Experiment Hayashi et al. (1989)
    CCSDc CCSD(T)d CCSDe  
A0   16578.647 16572.974 16557.528 16556.964(25)
B0   7344.446 7342.763 7334.765 7335.312(5)
C0   5519.346 5518.023 5511.869 5512.033(6)
ΔJ   8.124 8.404 7.865 8.02(11)
ΔJK   −34.249 −35.179 −31.683 −32.27(116)
ΔK   113.432 114.459 114.358 110.73(501)
δJ   2.917 3.031 2.771 2.78(7)
δK   2.267 2.291 2.567 2.35(138)
ΦJ   −0.0610      
ΦJK   0.0569      
ΦKJ   1.1229      
ΦK   −3.3337      
ϕJ   −0.0304      
ϕJK   −0.2534      
ϕK   0.5996      
s-CH332SCH2Dg
Parameter   Calculated Scaled Experiment Hayashi et al. (1989)
    CCSDc CCSD(T)d CCSDe  
A0   17819.084 17812.372 17796.684 17795.061(38)
B0   7077.495 7076.097 7068.186 7068.973(10)
C0   5408.647 5377.454 5401.313 5401.701(9)
ΔJ   6.504 6.721 6.297 6.87(21)
ΔJK   −31.194 −31.985 −31.448 −27.50(85)
ΔK   133.574 134.645 123.569 134.88(850)
δJ   2.234 2.321 2.123 2.02
δK   3.475 3.540 3.935  
ΦJ   −0.0480      
ΦJK   −0.1374      
ΦKJ   2.8026      
ΦK   −8.7476      
ϕJ   −0.0239      
ϕJK   −0.3493      
ϕK   0.6581      

Notes. aRotational constants in MHz; quartic centrifugal-distortion constants in kHz; sextic centrifugal-distortion constants in Hz. Watson A-reduction in the I r representation is used. bEquilibrium constants corresponding to the best-estimated equilibrium structure (see text; Senent et al. 2014). Vibrational corrections to rotational constants, and quartic and sextic centrifugal-distortion constants at the MP2/cc-pVTZ level. cEquilibrium constants corresponding to the best-estimated equilibrium structure (see text; Senent et al. 2014). Vibrational corrections to rotational constants, and quartic and sextic centrifugal-distortion constants at the CCSD/cc-pVTZ level. dEquilibrium constants corresponding to the best-estimated equilibrium structure (see text; Senent et al. 2014). Vibrational corrections to rotational constants, and quartic and sextic centrifugal-distortion constants at the CCSD(T)/cc-pVTZ level. eEmpirically scaled parameters (see text) starting from those computed as at footnote c. fDeuterium substitution leads to a complete symmetry loss: from the C2V symmetry point group to C1. gDeuterium substitution leads to a limited symmetry loss: from C2V to CS.

Download table as:  ASCIITypeset images: 1 2

Based on comparison with experiments, we note very good agreement between the latter and our best data. Indeed, we note mean discrepancies of about 0.12% and 0.07% when the best-estimated equilibrium constants (from the best-estimated equilibrium structures computed in Senent et al. 2014) are corrected for vibrational corrections at the CCSD/cc-pVTZ and CCSD(T)/cc-pVTZ levels, respectively. Employing MP2/cc-pVTZ corrections does not worsen the agreement, with mean discrepancies still on the order of 0.12%. Moving to quartic centrifugal-distortion constants, the MP2, CCSD, and CCSD(T) levels in conjunction with the cc-pVTZ basis set show averaged discrepancies of about 6.4%, 5.0%, and 6.1%, respectively. Sextic centrifugal-distortion constants deserve special note since, to our knowledge, only a few experimental values are available for the main isotopologue of DMS (Vacherand et al. 1987, see Table 1) and for the 32S- and 34S-containing isotopologues of gauche-ETSH and for 32S-trans-ETSH. While for DMS only a limited number of sextics are experimentally known, and therefore the poor agreement can be ascribed to limitations in the experimental work, for both gauche-and trans-ETSH the full sets of sextic centrifugal-distortion constants were determined in conjunction with higher-order terms. Therefore, a thorough inspection of this disagreement is required, especially because a sign inversion is observed in almost all cases. The corresponding discussion is reported later in the text.

Based on the results of Tables 13, we also note improved accuracy once we apply the empirical scaling procedure described above. In fact, the discrepancies decrease to less than 0.01% for rotational constants and to 3.1% (<2.0% in most cases) on average for quartic centrifugal-distortion terms. This improvement also has repercussions for the prediction of rotational transitions. In fact, our best-estimated computed parameters provide predictions with a relative accuracy of ∼0.1% in the centimeter/millimeter-wave region. Therefore, a rotational frequency of 200 GHz is currently predicted with an accuracy of 200 MHz. For the same frequency, the relative error decreases to ∼0.01%, resulting in an accuracy of 20 MHz for our prediction of a rotational transition at 200 GHz.

Our overall conclusion is that our predicted parameters, also the nonscaled ones, are reliable. We thus consider them to be sufficiently accurate for supporting laboratory or astronomical assignments and identifications. From the discussion above and from a methodological point of view, we would like to point out that the MP2/cc-pVTZ level is suitable for obtaining vibrational corrections to rotational constants and centrifugal-distortion parameters with an accuracy that allows for quantitative predictions.

Concerning the sextic centrifugal-distortion constants of gauche- and trans-ETSH, as mentioned above, an essentially complete disagreement is observed. For the trans form, the experimental parameters are affected by large uncertainties and we also note that the quartic term ΔK is enlarged as it moves from gauche to trans, while theory predicts the opposite trend. In the case of gauche-ETSH, however, the sextics are well determined. The sextics are very similar to our computed values but are opposite in sign. Based on the available literature for these topics (see, for example, Puzzarini et al. 2012a, 2012b, 2014b), sextic centrifugal-distortion constants computed at the CCSD(T)/cc-pVTZ level usually have an averaged accuracy of about 10% and show maximum discrepancies of about 20%. Therefore, we suggest that we further investigate the rotational spectra of gauche- and trans-ETSH.

3.2. Torsional Energy Levels and Splittings

Table 4 summarizes the lowest torsional energy levels (ground and first torsional states) corresponding to the various isotopic species of g-ETSH, t-ETSH, and DMS. Second torsional states, combination bands, and vibrational partition functions are also provided in our tables. The levels are classified using the representations of the G6 (ETSH) and G36 (DMS) Molecular Symmetry Groups and using two vibrational quanta. For ETSH, ν20 and ν21 refer to the methyl torsion and the hydroxyl torsion, respectively. For DMS, ν15 represents the infrared active mode. For both g-ETSH and t-ETSH, the vibrational ground state A/E splitting caused by the methyl internal rotation is very small. Nevertheless, the vibrational ground state of the hydrogenated isotopologues of g-ETSH splits by ∼0.060 cm−1 as a consequence of the SH torsion; this splitting largely reduces and becomes very small (only 0.002 cm−1) in the deuterated species. The vibrational ground state of DMS splits into nine components. The levy of degeneracy leads to energy separations of less than 0.001 cm−1. This means that the rotational study of the vibrational ground state of DMS using a treatment for semi-rigid molecules is realistic. For the first and second excited states, non-rigidity needs to be considered.

Table 4. Low Torsional Energy Levels (in cm−1) of Ethyl Mercaptan and Dimethyl Sulfide Isotopologues Calculated at the CCSD(T)/aug-cc-pVTZ Level

gauche- ethyl-mercaptan
    CH3CH232SH CH3CH234SH CH3CH236SH CH3CH233SH CH3CH232SD
ZPVE 226.898 226.773 226.661 226.834 201.658
ν20, ν21            
0 0+ A1, E 0.000 0.000 0.000 0.000 0.000
0 0 A2, E 0.061 0.060 0.059 0.060 0.002
0 1+ A1, E 188.132 188.086 188.043 188.108 145.434
0 1 A2, E 189.613 189.552 189.497 189.582 145.525
1 0+ A1 254.046 253.888 253.745 253.965 252.600
  E 254.045 253.887 253.746 253.964 252.600
1 0 A2 254.068 253.910 253.769 253.986 252.599
  E 254.067 253.909 253.768 253.985 252.598
0 2+ A1, E 339.848 339.824 339.802 339.835 275.486
0 2 A2, E 356.237 356.102 355.981 356.168 276.729
1 1+ A1 441.675 441.493 441.331 441.581 398.580
  E 441.677 491.496 441.334 441.584 398.581
1 1 A2 441.334 441.175 441.033 441.252 398.982
  E 441.336 441.178 441.036 441.255 398.983
2 0+ A1 491.359 491.049 490.772 491.199 487.356
  E 491.145 491.074 490.797 491.223 487.369
2 0 A2 491.120 490.804 490.521 490.957 487.343
  E 491.384 490.829 490.546 490.982 487.383
trans- ethyl-mercaptan
    CH3CH232SH CH3CH234SH CH3CH236SH CH3CH233SH CH3CH232SD
ν20, ν21            
0 0 A1, E 157.835 157.991 158.133 157.915 160.083
0 1 A2, E 313.572 313.669 313.760 313.622 283.577
1 0 A2 403.069 403.146 403.218 403.108 403.069
  E 403.067 403.145 403.217 403.107 403.068
0 2 A1, E 424.286 424.349 424.409 424.318 392.125
1 1 A1 555.804 555.706 555.622 555.753 522.955
  E 555.807 555.709 555.624 555.756 522.956
2 0 A1 632.303 632.332 632.362 632.317 628.722
  E 632.342 632.371 632.400 632.357 628.765
Dimethyl sulfide
      CH332SCH3 CH334SCH3 CH336SCH3 CH333SCH3
ZPVE 187.066 186.905 186.761 186.983
ν11, ν15            
0 0 A1 0.000 0.000 0.000 0.000
  G, E1, E3 0.001 0.001 0.001 0.001
1 0 A3 176.516 176.530 176.542 176.523
  G 176.500 176.514 176.526 176.507
  E2, E3 176.484 176.498 176.511 176.491
0 1 A2 182.289 181.983 181.709 182.131
  G 182.273 181.967 181.694 182.116
  E1, E4 182.258 181.952 181.678 182.100
2 0 A1 339.979 339.888 339.802 339.933
  G 340.213 340.114 340.018 340.163
  E1, E3 340.587 340.494 340.406 340.540
1 1 A4 341.312 341.058 340.831 341.181
  G 341.712 341.462 341.240 341.583
  E2, E4 341.972 341.711 341.478 341.837
0 2 A1 361.048 360.590 360.185 360.811
  G 361.044 360.584 360.177 360.806
  E1, E3 361.040 360.577 360.169 360.800

Download table as:  ASCIITypeset image

The variation of the torsional energies with the substitution of 32S for other sulfur isotopes is very small (less than 1 cm−1) because only the hydrogen atoms are significantly displaced during the internal rotation. Nevertheless, the variation is very important for the SH→SD substitution. For example, the OH fundamental levels vary from 188.132 cm−1 and 189.613 cm−1 to 145.434 cm−1 and 145.525 cm−1. The CH3 fundamentals vary from 254.046 (5) cm−1 and 254.068(7) cm−1 to 252.600 cm−1 and 252.599(8) cm−1. It must be considered that deuterium is a cosmological abundant isotope. CH3CH2SD represent the most probable detectable compound.

4. CONCLUDING REMARKS

The present contribution provides an accurate and reliable set of spectroscopic parameters for different isotopic species of DMS and ETSH. For this purpose, state-of-the-art computational methods and approaches have been employed. We thus determined highly accurate rotational and torsional parameters. Available experimental data were used to assess the reliability of our computations. For instance, relative accuracies of 0.1% and 5%–6% were observed for rotational and quartic centrifugal-distortion constants, respectively, and were further improved through empirical scaling. The rotational-spectroscopy characterization was complemented by the calculation of reliable sextic centrifugal-distortion terms. On the whole, our computed parameters allow us to predict rotational transitions with the proper accuracy for future laboratory and/or astronomical investigations.

The A/E splitting of the ground vibrational state of ETSH caused by the A/E torsion is almost negligible. However, a splitting of ∼0.060 cm−1 due to the SH torsion is observed for the hydrogenated isotopologues of g-ETSH; this splitting is negligible in the deuterated species. In DMS, the low-energy torsional states split into nine components with separations of less than 0.001 cm−1. Therefore, the vibrational ground-state rotational study based on the semi-rigid rotor approximation should be considered reliable; meanwhile, an appropriate treatment accounting for non-rigidity is required to describe the first and second excited states.

The overall conclusion is that based on the good agreement and well-established computational techniques employed, we are confident that the spectroscopic data provided herein are highly accurate and can therefore be useful for the identification of rare isotopologues of DMS and ETSH in the interstellar medium.

This research was supported by a Marie Curie International Research Staff Exchange Scheme Fellowship within the 7th European Community Framework Program under grant No. PIRSES-GA-2012‐31754, the COST Action CM1002 CODECS, and the FIS2011‐28738-C02‐02 project (MINECO, Spain). In Bologna, this work was supported by MIUR (PRIN 2012 funds under the project "STAR: Spectroscopic and computational Techniques for Astrophysical and atmospheric Research") and by the University of Bologna (RFO funds). M.L.S., M.H., and M.A.M. acknowledge the Deanship of Scientific Research at King Saud University for its funding through the Research Group RGP-VPP-333.

The authors acknowledge CTI (CSIC) and CESGA for computing facilities.

Please wait… references are loading.
10.1088/0004-637X/796/1/50