on the intermolecular interactions of isothiocyanic acid (hncs) with disulfur monoxide (sso): a...

11
This article was downloaded by: [UOV University of Oviedo] On: 28 October 2014, At: 11:45 Publisher: Taylor & Francis Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK Journal of Sulfur Chemistry Publication details, including instructions for authors and subscription information: http://www.tandfonline.com/loi/gsrp20 On the intermolecular interactions of isothiocyanic acid (HNCS) with disulfur monoxide (SSO): a first principles approach Esmail Vessally a , Ali Mortezapour b & Moein Goodarzi c a Department of Science, Payame Noor University, P. O. Box 19395–4697, Tehran, Iran b Department of Physics, University of Guilan, P. O. Box 41335–19141, Rasht, Iran c Department of Chemistry, Institute for Advanced Studies in Basic Sciences (IASBS), Zanjan, Iran Published online: 23 May 2014. To cite this article: Esmail Vessally, Ali Mortezapour & Moein Goodarzi (2014) On the intermolecular interactions of isothiocyanic acid (HNCS) with disulfur monoxide (SSO): a first principles approach, Journal of Sulfur Chemistry, 35:5, 484-492, DOI: 10.1080/17415993.2014.917376 To link to this article: http://dx.doi.org/10.1080/17415993.2014.917376 PLEASE SCROLL DOWN FOR ARTICLE Taylor & Francis makes every effort to ensure the accuracy of all the information (the “Content”) contained in the publications on our platform. However, Taylor & Francis, our agents, and our licensors make no representations or warranties whatsoever as to the accuracy, completeness, or suitability for any purpose of the Content. Any opinions and views expressed in this publication are the opinions and views of the authors, and are not the views of or endorsed by Taylor & Francis. The accuracy of the Content should not be relied upon and should be independently verified with primary sources of information. Taylor and Francis shall not be liable for any losses, actions, claims, proceedings, demands, costs, expenses, damages, and other liabilities whatsoever or howsoever caused arising directly or indirectly in connection with, in relation to or arising out of the use of the Content. This article may be used for research, teaching, and private study purposes. Any substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any form to anyone is expressly forbidden. Terms &

Upload: moein

Post on 04-Mar-2017

223 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: On the intermolecular interactions of isothiocyanic acid (HNCS) with disulfur monoxide (SSO): a first principles approach

This article was downloaded by: [UOV University of Oviedo]On: 28 October 2014, At: 11:45Publisher: Taylor & FrancisInforma Ltd Registered in England and Wales Registered Number: 1072954 Registeredoffice: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Journal of Sulfur ChemistryPublication details, including instructions for authors andsubscription information:http://www.tandfonline.com/loi/gsrp20

On the intermolecular interactions ofisothiocyanic acid (HNCS) with disulfurmonoxide (SSO): a first principlesapproachEsmail Vessallya, Ali Mortezapourb & Moein Goodarzica Department of Science, Payame Noor University, P. O. Box19395–4697, Tehran, Iranb Department of Physics, University of Guilan, P. O. Box41335–19141, Rasht, Iranc Department of Chemistry, Institute for Advanced Studies in BasicSciences (IASBS), Zanjan, IranPublished online: 23 May 2014.

To cite this article: Esmail Vessally, Ali Mortezapour & Moein Goodarzi (2014) On the intermolecularinteractions of isothiocyanic acid (HNCS) with disulfur monoxide (SSO): a first principles approach,Journal of Sulfur Chemistry, 35:5, 484-492, DOI: 10.1080/17415993.2014.917376

To link to this article: http://dx.doi.org/10.1080/17415993.2014.917376

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the“Content”) contained in the publications on our platform. However, Taylor & Francis,our agents, and our licensors make no representations or warranties whatsoever as tothe accuracy, completeness, or suitability for any purpose of the Content. Any opinionsand views expressed in this publication are the opinions and views of the authors,and are not the views of or endorsed by Taylor & Francis. The accuracy of the Contentshould not be relied upon and should be independently verified with primary sourcesof information. Taylor and Francis shall not be liable for any losses, actions, claims,proceedings, demands, costs, expenses, damages, and other liabilities whatsoever orhowsoever caused arising directly or indirectly in connection with, in relation to or arisingout of the use of the Content.

This article may be used for research, teaching, and private study purposes. Anysubstantial or systematic reproduction, redistribution, reselling, loan, sub-licensing,systematic supply, or distribution in any form to anyone is expressly forbidden. Terms &

Page 2: On the intermolecular interactions of isothiocyanic acid (HNCS) with disulfur monoxide (SSO): a first principles approach

Conditions of access and use can be found at http://www.tandfonline.com/page/terms-and-conditions

Dow

nloa

ded

by [

UO

V U

nive

rsity

of

Ovi

edo]

at 1

1:45

28

Oct

ober

201

4

Page 3: On the intermolecular interactions of isothiocyanic acid (HNCS) with disulfur monoxide (SSO): a first principles approach

Journal of Sulfur Chemistry, 2014Vol. 35, No. 5, 484–492, http://dx.doi.org/10.1080/17415993.2014.917376

On the intermolecular interactions of isothiocyanic acid (HNCS)with disulfur monoxide (SSO): a first principles approach

Esmail Vessallya∗, Ali Mortezapourb and Moein Goodarzic∗

aDepartment of Science, Payame Noor University, P. O. Box 19395–4697, Tehran, Iran; bDepartment ofPhysics, University of Guilan, P. O. Box 41335–19141, Rasht, Iran; cDepartment of Chemistry, Institute forAdvanced Studies in Basic Sciences (IASBS), Zanjan, Iran

(Received 16 January 2014; accepted 18 April 2014 )

The geometrical structure and binding energy of all the possible heterodimers of isothiocyanic acid (HNCS)with disulfur monoxide (SSO) have been studied in the gas phase, theoretically. Nine minima (with noimaginary frequencies) are located on the singlet potential energy surface of the HNCS–SSO system at theMP2 level with binding energies (corrected with ZPE and BSSE) in the range 5.53–19.12 kJ/mol. Bader’squantum theory of atoms in molecules has been employed to elucidate the intermolecular interactioncharacteristics of the HNCS–SSO system. All intermolecular interactions in the HNCS–SSO system areweak interactions of non-covalent without any covalent characters.

Keywords: isothiocyanic acid; disulfur monoxide; intermolecular interaction; binding energy; AIM

1. Introduction

The isothiocyanic acid (HNCS) is a sulfur analog of the well-known isocyanic acid (HNCO).Many investigations in both of theoretical and experimental have been performed on the structure,

∗Corresponding author. Emails: [email protected]; [email protected] article makes reference to supplementary material available on the publisher’s website at http://dx.doi.org/10.1080/17415993.2014.917376.

© 2014 Taylor & Francis

Dow

nloa

ded

by [

UO

V U

nive

rsity

of

Ovi

edo]

at 1

1:45

28

Oct

ober

201

4

Page 4: On the intermolecular interactions of isothiocyanic acid (HNCS) with disulfur monoxide (SSO): a first principles approach

Journal of Sulfur Chemistry 485

stability and photolytic behavior of HNCO and its structural isomers.[1–12] In contrast, HNCSand its isomers were less studied up to now. There are several papers on its IR,[13,14] microwave[15,16] and UV [17] spectra. Flash photolysis of HNCS was also reported.[18] Durig and Wertz[19] have performed detailed infrared investigation of HNCS and DNCS in gaseous and solidphases and in argon matrices. High resolution rotational–vibrational spectra of the acid have alsobeen studied.[20] Later on, theoretical [21] and the rotational spectrum of HNCS [22,23] haveconfirmed the structure of the HNCS as a quasi-linear molecule. Wierzejewska and Mielke [24]have reported near UV photolysis and the infrared matrix isolation studies of HNCS.[25]Also, theyhave carried out ab initio calculations on HNCS and its complexes with nitrogen and xenon.[25]The theoretical study of Bak et al. [21] shows that a stability of HNCS and its isomers should be inthe order of HNCS > HSCN > HSNC > HCNS. Durig et al. [26] have determined the structuralparameters and vibrational spectra of the HNCS molecule, theoretically. Very recently, Wang et al.[27] have investigated the mechanism of thermal decomposition of thiourea, theoretically. Theirresults show that the HNCS can be easily formed through thermal decomposition of thiourea.

One of the interesting molecules which is predicted to be the component of atmosphere of theJupiter’s moon Io [28,29] and the planet Venus [30] is the disulfur monoxide (SSO) molecule.Also, the SSO molecule is reported to be produced in the earth’s atmosphere by the photo-chemical oxidation reactions of hydrogen disulfide and hydrocarbon disulfide compounds. Thesereactions suggest that SSO would be an intermediate from the photochemical oxidation of disul-fides. Because of the small amount of production or interference of by-products, the most of thesereactions are not suitable for experimentally investigating the characteristics of SSO. A relativelyclean process is the reaction of oxygen atoms with OCS, producing SO and CO. The decay ofSO, in the absence of O2 has been studied at 298 K from 2 to 8 Torr using a tubular flow reactorcoupled to a mass spectrometer. According to this study the formation of SSO from gaseousSO is a stepwise process.[31] Therefore, the reaction of oxygen atoms with OCS is suitable toproduce the SSO molecule. The formation of SSO is also reported to occur from the gaseousreactions of SO in a sulfur-rich environment.[32] The structure of the SSO molecule has beenreported in many published research papers.[33–40] Experimentally, the structure of the SSOmolecule has been determined by rotational spectroscopy.[41–43] Theoretically, high-level abinitio calculations support the open-chain structure for the ground state of SSO,[32,44] whiledensity functional calculations predict that a cyclic ground state structure is more stable.[45] Thereaction paths of SSO with O2 have been investigated theoretically, where the molecules of SO2

and SO are the products of the SSO + O2 reaction.[46] This means that the SSO molecule affectsacid rain because the formed SO reacts with O2 to produce SO3.[46–48]

Since the latter species play an important role in the earth’s atmosphere, it is important toobtain atmospheric information regarding their behaviors. In the present work, we study theheterodimers containing intermolecular interactions of HNCS with the SSO molecule to providea detailed examination of the stability and property of the HNCS–SSO system.

2. Computational methods

Ab initio calculations are carried out using the Gaussian 03 programs.[49] The geometries ofHNCS and SSO monomers and all the heterodimers are fully optimized by employing the Moller–Plesset second-order perturbation (MP2) method [50] with Dunning’s correlation-consistent basisset, cc-pVDZ [51] on the singlet potential energy surface (PES). In order to compute accu-rate relative energies, single-point energy calculations at the MP2/aug-cc-pVTZ level have beencarried out on the optimized geometries at the MP2/cc-pVDZ computational level (MP2/aug-cc-pVTZ//MP2/cc-pVDZ). Note that the interaction energy (EI) was calculated as the differencebetween the total energy of the heterodimer and the sum of total energies of the HNCS and SSO

Dow

nloa

ded

by [

UO

V U

nive

rsity

of

Ovi

edo]

at 1

1:45

28

Oct

ober

201

4

Page 5: On the intermolecular interactions of isothiocyanic acid (HNCS) with disulfur monoxide (SSO): a first principles approach

486 E. Vessally et al.

monomers. The vibrational frequencies are calculated at the MP2/cc-pVDZ level of the theorybased on the optimized geometries to determine the nature of the stationary points according tothe number of negative eigenvalues of the Hessian matrix. The heterodimers possess no imaginaryfrequencies. Also, the calculation of the vibrational frequencies leads to the evaluation of zero-point energies (ZPEs). The counterpoise (CP) method [52] has been used to take into account thebasis set superposition error (BSSE) in the calculation of the energies and presents the energieswithout any BSSEs. Finally, the quantum theory of atoms in the molecules analysis [53] wasperformed with the help of the AIM 2000 software [54] using the wave functionals generated atthe MP2/cc-pVDZ level.

3. Results and discussion

The electronic structures of the isolated HNCS and SSO monomers were used to generate a setof heterodimers on the singlet PES. Note that the symbol A stands for the heterodimers of theHNCS–SSO system and the numbering of the symbolA conforms to the ordering of binding energy(corrected with BSSE and ZPE at the MP2/aug-cc-pVTZ//MP2/cc-pVDZ level). The optimizedgeometries of all the heterodimers and their ranking are reported in Figures 1 and 2, respectively.Also, the calculated binding energies and bond critical point data for different heterodimers in theMP2 level were listed in Tables 1 and 2, respectively. Note that the standard orientations of all het-erodimers of the HNCS–SSO system at the MP2/cc-pVDZ level and more details of energies are inTables S1 and S2 of supporting information (http://dx.doi.org/10.1080/17415993.2014.917376).

Association of the HNCS and the SSO monomers leads to the formation of nine minima, A1throughA9, on the singlet PES. Binding energies of the HNCS–SSO heterodimers including BSSEand ZPE corrections lie in the range of 5.23–19.12 kJ/mol at the MP2/aug-cc-pVTZ//MP2/cc-pVDZ level.

Figure 1. Optimized geometries of all the heterodimers in the HNCS–SSO system at the MP2/cc-pVDZ level (the bondlengths are in angstrom).

Dow

nloa

ded

by [

UO

V U

nive

rsity

of

Ovi

edo]

at 1

1:45

28

Oct

ober

201

4

Page 6: On the intermolecular interactions of isothiocyanic acid (HNCS) with disulfur monoxide (SSO): a first principles approach

Journal of Sulfur Chemistry 487

Figure 2. The ranking of the heterodimers obtained in the HNCS–SSO system in terms of EI + BSSE + ZPE at theMP2/aug-cc-pVTZ//MP2/cc-pVDZ level.

Table 1. Binding energies (kJ/mol) of heterodimers obtained in theHNCS–SSO system at the MP2 level.

Optimization Single-point energySpecies MP2/cc-pVDZ MP2/aug-cc-pVTZ//MP2/cc-pVDZ

EI EI EI + BSSE EI + BSSE + ZPE

A1 −34.24 −28.68 −24.46 −19.12A2 −13.92 −20.80 −17.91 −15.68A3 −26.96 −22.33 −18.62 −14.16A4 −11.42 −15.79 −12.98 −10.43A5 −14.55 −16.52 −13.21 −10.24A6 −13.63 −14.92 −11.69 −8.83A7 −10.45 −10.35 −8.17 −6.17A8 −10.35 −10.19 −8.01 −6.09A9 −9.35 −8.41 −6.59 −5.23

A1 is the most stable heterodimer of the HNCS–SSO system with a binding energy 19.12 kJ/molafter corrections of BSSE and ZPE at the MP2/aug-cc-pVTZ//MP2/cc-pVDZ level. Figure 1shows that the planar heterodimer of A1 forms when O- and S3-atoms of the SSO monomer

Dow

nloa

ded

by [

UO

V U

nive

rsity

of

Ovi

edo]

at 1

1:45

28

Oct

ober

201

4

Page 7: On the intermolecular interactions of isothiocyanic acid (HNCS) with disulfur monoxide (SSO): a first principles approach

488 E. Vessally et al.

Table 2. Interatomic distances (Å) and bond critical point data (a.u.) calculated at theMP2/cc-pVDZ level.

Heterodimers Interaction Interatomic distance ρ ∇2ρ −GC/VC

A1 O…H 1.890 0.0256 0.1034 1.1602S3…N 3.806 0.0043 0.0129 1.2517

A2 S2…S1 3.602 0.0073 0.0188 1.2994S3…N 3.300 0.0091 0.0244 1.1397

A3 O…H 1.897 0.0246 0.1003 1.1684A4 S2…N 3.126 0.0103 0.0312 1.1978

O…C 3.204 0.0067 0.0218 1.2048A5 S3…H 2.626 0.0119 0.0309 1.0967A6 S3…H 2.596 0.0124 0.0327 1.1025A7 O…S1 3.135 0.0084 0.0312 1.1769

S3…S1 3.930 0.0045 0.0114 1.1815A8 O…S1 3.132 0.0085 0.0314 1.1765

S3…S1 3.965 0.0043 0.0111 1.1801A9 O…S1 3.212 0.0074 0.0280 1.2035

attack H- and N-atoms of the HNCS monomer to create the interactions of O…H and S3…N,respectively. The bond lengths of the interactions of O…H and S3…N are 1.890 and 3.806Å atthe MP2/cc-pVDZ level, respectively.

As shown in Table 2, binding energy of the A2 heterodimer is only 3.44 kJ/mol less than that ofthe A1 heterodimer at the MP2/aug-cc-pVTZ//MP2/cc-pVDZ level. The atomic connectivity ofthe A2 heterodimer is similar to the A1. Both of A1 and A2 heterodimers include intermolecularinteraction of S3…N. Their main difference is the second interaction. In the A2 heterodimer, theS2-atom of the SSO reacts with the Sl-atom of the HNCS monomer (S2…S1 interaction), whilein the A1 heterodimer, the O-atom of the SSO attacks the H-atom of the HNCS monomer (O…Hinteraction). This means that the replacement of the S2…S1 interaction in the A2 heterodimerwith the O…H interaction creates a more stable heterodimer of the A1.

A3 heterodimer includes only one interaction between the SSO and the HNCS monomers onPES. In theA3 heterodimer, O-atom of the SSO approaches H-atom of the HNCS monomer to formO…H interaction with bond length 1.897Å. This heterodimer has binding energy 14.16 kJ/mol(corrected with BSSE and ZPE) at the MP2/aug-cc-pVTZ//MP2/cc-pVDZ level. Note that theO…H interaction of the A3 heterodimer is slightly longer than that of the A1 as given in Table 2.

A4 heterodimer has the fourth stability order among nine heterodimers obtained for the HNCS–SSO system. It lies 10.43 kJ/mol below the initial monomers of the HNCS and the SSO. It isinteresting to say that two interactions of O…C (3.204Å) and S2…N (3.126Å) have been observedonly in the A4 heterodimer.

The comparison of theA5 andA6 heterodimers shows that atomic connectivity of them is equal.In both of them, the S3-atom of the SSO monomer reacts with the H-atom of the HNCS monomerto produce the S3…H interaction. In spite of similar atomic connectivity, the A5 heterodimer(10.24 kJ/mol) is 1.41 kJ/mol stable than A6 (8.83 kJ/mol) at the MP2/aug-cc-pVTZ//MP2/cc-pVDZ level. The calculated results include that the S…H interaction exists only in two the A5and A6 heterodimers.

Similar to the A5 and the A6 heterodimers, the atomic connectivity of the A7 heterodimer isexactly similar to that of A8. Both of them include two interactions of O…S1 and S3…S1 asshown in Figure 1. Precise investigation of the situation of the HNCS H-atom relative to the SSOmonomer in the A7 and A8 heterodimers reveals that it is suitable to use trans and cis terms forthe A7 and the A8 heterodimers, respectively. The binding energies of the A7 and the A8 are 6.17and 6.09 kJ/mol at the MP2/aug-cc-pVTZ//MP2/cc-pVDZ level.

Dow

nloa

ded

by [

UO

V U

nive

rsity

of

Ovi

edo]

at 1

1:45

28

Oct

ober

201

4

Page 8: On the intermolecular interactions of isothiocyanic acid (HNCS) with disulfur monoxide (SSO): a first principles approach

Journal of Sulfur Chemistry 489

A9 heterodimer is the least stable heterodimer of the HNCS–SSO system with the bindingenergy 5.23 kJ/mol (corrected with BSSE and ZPE). The attack of the SSO terminal atom on theS1-atom of the HNCS monomer leads to the O…S1 interaction with the bond length 3.212Å inthe A9 heterodimer.

Wierzejewska and Wieczorek [25] studied the intermolecular interactions of the HNCS–N2

and HNCS–Xe systems at the MP2 level. They reported the existence of two and one stable com-plexes for the HNCS–N2 and HNCS–Xe systems, respectively. The comparison of our resultswith those of Wierzejewska et al. includes three important points. Firstly, the number of het-erodimers reported for the HNCS–SSO system in the present work (nine heterodimers) is muchmore than those of Wierzejewska et al. for the HNCS–N2 (two heterodimers) and HNCS–Xe(one heterodimer) systems. Secondly, the binding energy of the most stable heterodimer (A1) ofthe HNCS–SSO system corrected with only BSSE (24.46 kJ/mol) is much more than that of theHNCS–N2 (6.68 kJ/mol) and the HNCS–Xe (3.64 kJ/mol) systems. This means that the HNCS–SSO system has more important influence than the HNCS–N2 and the HNCS–Xe systems in theatmosphere. Thirdly, the most stable heterodimers obtained for two systems of the HNCS–N2

and the HNCS–Xe [25] involve an interaction from HN group of the HNCS molecule with eitherthe N2 molecule (in the HNCS–N2 system) or Xe (in the HNCS–Xe system). This is in goodagreement with intermolecular interactions obtained for the A1 heterodimers as the most stableheterodimers of the HNCS–SSO system in the present work.

As shown in Figure 2, the binding energies corrected with BSSE and ZPE for all the het-erodimers obtained in the HNCS–SSO system have been ranked to provide overall insight intobetween the binding energies and intermolecular interaction. It can be seen that all three unstableheterodimers of the HNCS–SSO system (A7, A8 and A9) include only an interactions betweenthe terminal sulfur atom of the HNCS monomer (denoted as S1) with the SSO monomer. Also,three heterodimers of A3, A5 and A6 have only one interaction between the HNCS and the SSOmonomers. As shown in Figure 2, the H-atom interaction of the HNCS with the O-atom of theSSO monomer (in the A3 heterodimer) creates a more stable complex relative to the H-atominteraction of the HNCS with the S-atom of the SSO monomer (in the A5 and A6 heterodimers).

It is necessary to say that all the possible intermolecular interactions in the of the HNCS–SSOsystem have been found in the present work except four interactions of O…N, S2…H, S2…C andS3…C. This means that none of the HNCS–SSO heterodimers include an interaction betweeneither the S-atom of the SSO (S2 or S3) and the C-atom of the HNCS monomer or the O-atom ofthe SSO and the N-atom of the HNCS monomer.

3.1. AIM analysis

TheAIM theory [53] has been used in order to analyze the electron density (ρ) of the heterodimers.According to this analysis, ρ is used to describe the strength of a bond. In general, the larger thevalue of ρ is, the stronger the bond is. The Laplacian of the electron density (∇2ρ) describes thecharacteristic of the bond. Where ∇2ρ < 0 the bond is a covalent bond, as ∇2ρ > 0, the bondbelongs to the ionic bond, hydrogen bond and van der Waals interaction. The GC/VC ratio, beingGC and VC, respectively, the kinetic and potential energy density at bond critical point (BCP), wasused as a measure of the covalency in non-covalent interactions. Values greater than 1 generallyindicate a non-covalent interaction without covalent character while ratios smaller than unity areindicative of the covalent nature of the interaction.[55,56] The values of topological parameters(ρ, ∇2ρ and GC/VC) for each intermolecular BCPs between the HNCS and the SSO monomersare given in Table 2.

As shown in Table 2, ρ values are in the range of 0.0043–0.0256 a.u. It is interesting to knowthat the biggest value of ρ between the intermolecular interactions of the HNCS–SSO system

Dow

nloa

ded

by [

UO

V U

nive

rsity

of

Ovi

edo]

at 1

1:45

28

Oct

ober

201

4

Page 9: On the intermolecular interactions of isothiocyanic acid (HNCS) with disulfur monoxide (SSO): a first principles approach

490 E. Vessally et al.

belongs to the O…H interaction in the A1 heterodimer (0.0256 a.u.). While the smallest ρ valuebelongs to the S3…N (in the A1 with 0.0043 a.u.) and S3…S1 (in the A8 with 0.0043 a.u.)interactions. The investigation of values obtained for ∇2ρ in the HNCS–SSO system shows thatall of them are positive and should be within the range of 0.0111–0.1034 a.u. All values reportedfor the −GC/VC ratio of heterodimers in the HNCS–SSO system is greater than 1, as given inTable 2. Therefore, the small electron density and positive values of ∇2ρ with −GC/VC > 1 areindicating that all intermolecular interactions in the HNCS–SSO system are weak interactions ofnon-covalent without any covalent characters.

4. Conclusions

The present work reports a study of the structural and electronic properties of heterodimersbetween the HNCS and the SSO monomers on the singlet PES, theoretically. The calculatedresults show that the attack of the HNCS on the SSO monomer leads to formation of nine min-ima, A1 through A9. Their binding energies (corrected with BSSE and ZPE) lie in the range of5.23–19.12 kJ/mol at the MP2/aug-cc-pVTZ//MP2/cc-pVDZ level. All possible intermolecularinteractions in the HNCS–SSO system have been found in the present work except four interac-tions of O…N, S2…H, S2…C and S3…C. The calculated results reveal that the heterodimerscontaining only interactions between the terminal sulfur atom of the HNCS monomer (denotedas S1) with the SSO monomer (A7, A8 and A9) have the least binding energy in the HNCS–SSOsystem. Also, the binding energy of the individual interaction of the HNCS H-atom with theO-atom of the SSO monomer (in the A3 heterodimer) is bigger than the individual interaction ofthe HNCS H-atom with the S-atom of the SSO monomer (in the A5 and A6 heterodimers). Thecomparison of the binding energies of the HNCS–SSO system (the present work) with those ofHNCS–N2 and the HNCS–Xe systems (Wierzejewska et al.) shows that the heterodimers of theHNCS–SSO system are much more stable that the heterodimers obtained in both of the HNCS–N2

and the HNCS–Xe systems. Based on the AIM analysis, the small electron density, positive valuesof ∇2ρ and −GC/VC > 1 indicate that all intermolecular interactions in the HNCS–SSO systemare weak interactions of non-covalent without any covalent characters.

Supplemental data

Supplemental data for this article can be accessed at 10.1080/17415993.2014.917376.

References

[1] Jacox ME, Milligan DE. Low temperature infrared study of intermediates in the photolysis of HNCO and DNCO.J Chem Phys. 1964;40:2457–2460.

[2] Okabe H. Photodissociation of HNCO in the Vacuum Ultraviolet; Production of NCO A2� and NH (A3�, �c1).J Chem Phys. 1970;53:3507–3515.

[3] Bondybey VE, English JH, Mathews CW, Contolini RJ. Infrared spectra and isomerization of CHNO species in raregas matrices. J Mol Spectrosc. 1982;92:431–442.

[4] Spiglanin TA, Perry RA, Chandler DW. Internal state distributions of CO from HNCO photodissociation. J ChemPhys. 1987;87:1568–1576.

[5] Ruscic B, Berkowitz J. The H-NCO bond energy and �H0f (NCO) from photoionization mass spectrometric studiesof HNCO and NCO. J Chem Phys. 1994;100:4498–4508.

[6] Brown SS, Berghout HL, Crim FF. Internal energy distribution of the NCO fragment from near-threshold photolysisof isocyanic acid, HNCO. J Phys Chem. 1996;100:7948–7955.

[7] Klossika J-J, Flöethmann H, Beck C, Schinke R, Yamashita K. The topography of the HNCO(S1) potential energysurface and its implications for photodissociation dynamics. Chem Phys Lett. 1997;276:325–333.

Dow

nloa

ded

by [

UO

V U

nive

rsity

of

Ovi

edo]

at 1

1:45

28

Oct

ober

201

4

Page 10: On the intermolecular interactions of isothiocyanic acid (HNCS) with disulfur monoxide (SSO): a first principles approach

Journal of Sulfur Chemistry 491

[8] Klossika J-J, Flöethmann H, Beck C, Schinke R, Bittererová M. On the S1 → S0 internal conversion in thephotodissociation of HNCO. Chem Phys Lett. 1999;314:182–188.

[9] Coffey MJ, Berghout HL, Woods III E, Crim FF. Vibrational spectroscopy and intramolecular energy transfer inisocyanic acid (HNCO). J Chem Phys. 1999;110:10850–10862.

[10] Pettersson M, Khriachtchev L, Jolkkonen S, Räsänen M. Photochemistry of HNCO in solid Xe: Channels of UVphotolysis and creation of H2NCO radicals. J Phys Chem A. 1999;103:9154–9162.

[11] Kaledin AL, Cui Q, Heaven MC, Morokuma K. Ab initio theoretical studies on photodissociation of HNCOupon S1(

1A′′) ← S0(1A′) excitation: The role of internal conversion and intersystem crossing. J Chem Phys.

1999;111:5004–5016.[12] Pettersson M, Khriachtchev L, Lundell J, Jolkkonen S, Räsänen M. Photochemistry of HNCO in solid xenon:

photoinduced and thermally activated formation of HXeNCO. J Phys Chem A. 2000;104:3579–3583.[13] Reid C. The infrared absorption spectrum of isothiocyanic acid (HNCS). J Chem Phys. 1950;18:1512–1513.[14] Barakat TM, Legge N, Pullin ADE. Spectra and hydrogen-bonding characteristics of thiocyanic acid. Part 2. Spectra

of thiocyanic acid-base mixtures in inert solvents. Trans Faraday Soc. 1963;59:1773–1783.[15] Kewley R, Sastry KVLN, Winnewisser M. The millimeter wave spectra of isocyanic and isothiocyanic acids. J Mol

Spectrosc. 1963;10:418–441.[16] Beard CI, Dailey BP. Microwave spectrum and structure of isothiocyanic acid. J Chem Phys. 1947;15;762.[17] McDonald JR, Scherr VM, McGlynn SP. Lower-energy electronic states of HNCS, NCS−, and thiocyanate salts.

J Chem Phys. 1969;51:1723–1731.[18] Boxall CR, Simons JP. The photodissociation of HNCS in the near ultra-violet. J Photochem. 1972/1973;1:363–369.[19] Durig JR, Wertz DW. On the infrared spectra of HNCS and DNCS. J Chem Phys. 1967;46:3069–3077.[20] Draper GR, Werner RL. The rotational-vibrational spectra of HNCS and DNCS: An analysis of the high resolution

spectra. J Mol Spectrosc. 1974;50:369–402.[21] Bak B, Christaiansen JJ, Nielsen OJ, Svanholt H. Comparable ab initio calculated energies of HCNS, CNSH, NCSH

and HNCS. Optimized geometries and dipole moments. Acta Chem Scand A. 1977;31:666–668.[22] Yamada K, Winnewisser M, Winnewisser G, Szalanski LB, Gerry MCL. Ground state spectroscopic constants of

isothiocyanic HNCS, from its microwave and millimeter wave spectra combined with far infrared data. J MolSpectrosc. 1979;78:189–202.

[23] Yamada K. Winnewisser M. Winnewisser G. Szalanski LB. Gerry MCL. Ground state spectroscopic constants ofH15NCS, HN13CS, and HNC34S, and the molecular structure of isothiocyanic acid. J Mol Spectrosc. 1980;79:295–313.

[24] Wierzejewska M, Mielke Z. Photolysis of isothiocyanic acid HNCS in low-temperature matrix. Infrared detectionof HSCN and HSNC isomers. Chem Phys Let. 2001;349:227–234.

[25] Wierzejewska, M, Wieczorek R. Infrared matrix isolation and ab initio studies on isothiocyanic acid HNCS and itscomplexes with nitrogen and xenon. Chem Phys. 2003;287:169–181.

[26] Durig JR, Zheng C, Deeb H. On the structural parameters and vibrational spectra of some XNCS and XSCN (X = H,F, Cl, Br) molecules. J Mol Struct. 2006;784:78–92.

[27] Wang ZD,Yoshida M, George B. Theoretical study on the thermal decomposition of thiourea. Comput Theor Chem.2013;1017:91–98.

[28] Hapke B. The surface of Io: A new model, Icarus. 1989;79:56–74.[29] Baklouti D, Schmit B, Brissaud O. S2O, polysulfuroxide and sulfur polymer on Io’s surface? Icarus. 2008;194:

647–659.[30] Na CY, Esposito LW. Is disulfur monoxide a second absorber on venus? Icarus. 1997;125:364–368.[31] Herron JT, Huie RE. Rate constants at 298 k for the reactions SO + SO + M → (SO)2 + M and SO + (SO)2 →

SO2 + S2O. Chem Phys Lett. 1980;76:322–324.[32] Fueno T, Buenker RJ. Electronic structures of the S2O and S3 isomers: an ab initio CI study. Theor Chim Acta.

1988:73:123–134.[33] Thorwirth S, Theulé P, Gottlieb CA, Müller HSP, McCarthy MC, Thaddeus P. Rotational spectroscopy of

S2O: Vibrational satellites, 33S isotopomers, and the sub-millimeter-wave spectrum. J Mol Struct. 2006;795:219–229.

[34] Ishii A, Nakabayashi M, Jin Y, Nakayama J. Oxidation of tetrathiolanes: isolation of a vic-disulfoxide, 5-(1-adamantyl)-5-tert-butyltetrathiolane 2,3-dioxide and its decomposition to the dithiirane 1-oxide and ‘S2O’.J Organomet Chem. 2000;611:127–135.

[35] Müller T, Vaccaro PH, Pérez-Bernal F, Lachello F. Algebraic approach for the calculation of polyatomic Franck–Condon factors: application to the vibronically resolved absorption spectrum of disulfur monoxide (S2O). ChemPhys Lett. 2000;329:271–282.

[36] Paulse CD, Poirier RA, Davis RW. The molecular structures and harmonic force fileds of SO2, S2O, and S2O2 fromcombined spectroscopic and SCF results. Chem Phys Lett. 1990;172:43–48.

[37] Norwood K, Ng CY. Photoion-photoelectron coincidence spectroscopy of the transient molecules SO and S2O.Chem Phys Lett. 1989;156:145–150.

[38] Clouthier DJ, Rutherford ML. Supersonic jet spectroscopy of S2O. Chem Phys. 1988;127:189–196.[39] Clouthier D. The visible absorption spectrum of S2O. J Mol Spectrosc. 1987;124:179–184.[40] Tsukiyama K, Kobayashi D, Obi K, Tanaka I. The ultraviolet bands of S2O studied by LIF and optical-optical

double-resonance spectroscopy. Chem Phys. 1984;84:337–343.[41] Meschi DJ, Myers RJ. The microwave spectrum, structure, and dipole moment of disulfur monoxide. J Mol Spectrosc.

1959;3:405–416.

Dow

nloa

ded

by [

UO

V U

nive

rsity

of

Ovi

edo]

at 1

1:45

28

Oct

ober

201

4

Page 11: On the intermolecular interactions of isothiocyanic acid (HNCS) with disulfur monoxide (SSO): a first principles approach

492 E. Vessally et al.

[42] Tiemann E, Hoeft J, Lovas FJ, Johnson DR. Spectroscopic studies of the SO2 discharge system. I. The microwavespectrum and structure of S2O. J Chem Phys. 1974;60:5000–5004.

[43] Marsden CJ, Smith BJ. The harmonic force field and rz molecular structure of disulphur monoxide, S2O. ChemPhys. 1990;141:325–334.

[44] Ivanic J, Atchity GJ, Ruedenberg K. Violation of the weak noncrossing rule between totally symmetric closed-shellstates in the valence-isoelectronic series O3, S3, SO2, and S2O. J Chem Phys. 1997;107:4307–4317.

[45] Jones RO. Electronic structure of sos - a comparison with O3, SO2, and S3. Chem Phys Lett. 1986;125:221–224.

[46] Goodarzi M, Vahedpour M, Nazar F. Theoretical study on the atmosphere formation of SOx (x = 1–3) in the SSO(1A′) and O2 (3�−

g ) reaction. J Mol Struct (Theochem). 2010;945:45–52.[47] Li X, Fan H, Meng L, ZengY, Zheng S. Theoretical investigation on stability and isomerizations of CH3SO isomers.

J Phys Chem A. 2007;111:2343–2350.[48] Kim J, Lee K, Yang E, Shin K, Noh J, Park K, Hyun B, Jeong HJ, Kim JY, Kim K, Kim M, Kim H, Jang P, Jang

MC. Enhanced production of oceanic dimethylsulfide resulting from CO2-Induced grazing activity in a high CO2world. Environ Sci Technol. 2010;44:8140–8143.

[49] Frisch MJ, Trucks GW, Schlegel HB, Scuseria GE, Robb MA, Cheeseman JR, Montgomery JA Jr, Vreven T, KudinKN, Burant JC, Millam JM, Iyengar SS, Tomasi J, Barone V, Mennucci B, Cossi M, Scalmani G, Rega N, PeterssonGA, Nakatsuji H, Hada M, Ehara M, Toyota K, Fukuda R, Hasegawa J, Ishida M, Nakajima T, Honda Y, Kitao O,Nakai H, Klene M, Li X, Knox JE, Hratchian HP, Cross JB,Adamo C, Jaramillo J, Gomperts R, Stratmann RE,YazyevO, Austin AJ, Cammi R, Pomelli C, Ochterski JW, Ayala PY, Morokuma K, Voth GA, Salvador P, Dannenberg JJ,Zakrzewski VG, Dapprich S, Daniels AD, Strain MC, Farkas O, Malick DK, Rabuck AD, Raghavachari K, ForesmanJB, Ortiz JV, Cui Q, Baboul AG, Clifford S, Cioslowski J, Stefanov BB, Liu G, Liashenko A, Piskorz P, KomaromiI, Martin RL, Fox DJ, Keith T, Al-Laham MA, Peng CY, Nanayakkara A, Challacombe M, Gill PMW, Johnson B,Chen W, Wong MW, Gonzalez C, Pople JA. Gaussian 03, Revision B03. Pittsburgh, PA: Gaussian, Inc.; 2003.

[50] Møller C, Plesset MS. Note on an approximation treatment for many-electron systems. Phys Rev. 1934;46:618–622.[51] Dunning Jr TH. Gaussian basis sets for use in correlated molecular calculations. I. The atoms boron through neon

and hydrogen. J Chem Phys. 1989;90:1007–1024.[52] Boys SF, Bernardi F. The calculation of small molecular interactions by the differences of separate total energies.

Some procedures with reduced errors. Mol Phys. 1970;19:553–566.[53] Bader RFW, Halpen J, Green MLH. The international series of monographs of chemistry. Oxford: Clarendon Press;

1990.[54] Biegler-Konig F, Schonbohm J. AIM2000 Program Package, Ver. 2.0. Bielefeld: University of Applied Sciences;

2002.[55] Ziókowski M, Grabowski SJ, Leszczynski J. Cooperativity in hydrogen-bonded interactions: Ab initio and “atoms

in molecules” analyses. J Phys Chem A. 2006;110:6514–6521.[56] Rozas I, Alkorta I, Elguero J. Behavior of ylides containing N, O, and C atoms as hydrogen bond acceptors. J Am

Chem Soc. 2000;122:11154–11161.

Dow

nloa

ded

by [

UO

V U

nive

rsity

of

Ovi

edo]

at 1

1:45

28

Oct

ober

201

4