Computational Study of Structural, Vibrational and

0 downloads 0 Views 770KB Size Report
bond is based on the association between p valence orbitals of M and the 3p ... ongoing from phosphorus to bismuth showed an increase of ionic character and it might explain ... structure considered as aligned (Bi4S6)n ribbons bound to each other by the ... The electron lone pair of the M atom is localized in hybrid orbital.
1

Computational Study of Structural, Vibrational and Electronic Properties of the Highly

2

Symmetric Molecules M4S6 (M= P, As, Sb, Bi).

3 4

E. Semidalas and A. Chrissanthopoulos

5 6

Laboratory of Inorganic Chemistry, Department of Chemistry, National and Kapodistrian University of

7

Athens, University campus, Zografou, GR-15771, Greece.

8 9

ABSTRACT

10 11

A systematic computational investigation of the structural, electronic and vibrational properties

12

of the group 15 sulfides M4S6 at Td symmetry was carried out. The performance of DFT and

13

MP2 theoretical methods was assessed compared to the high-level CCSD method. The M-S

14

bond is based on the association between p valence orbitals of M and the 3p of sulfur according

15

to the natural population analysis. Both polarizability and polarizability volume of the cage

16

molecules increase as the size of the atoms increases from P to Bi. A structural ‘relaxation’

17

ongoing from phosphorus to bismuth showed an increase of ionic character and it might explain

18

the chemical instability of the heavier cage compounds. For the P4S6 molecule, the functionals

19

wB97XD and CAMB3LYP yielded excellent structural data, while for the heavier molecules

20

As4S6, Sb4S6 and Bi4S6, the M06 and M06L functionals showed high accuracy. We validated

21

eight functionals BP86, M06L, B3LYP, M06, Μ06-2Χ, CAMB3LYP, wB97XD, B2PLYP

22

which span from conventional GGA functionals to long-range corrected hybrid ones, and MP2,

23

CCSD ab initio methods. Experimentally, these molecules could be useful in the structural

24

investigation of the isolated gas phase species, besides solving complex structures of liquid,

25

crystalline or amorphous phases.

26 27 28

Keywords: M4S6 Td molecules, ab-initio, DFT, structural properties, electronic properties, vibrational

29

properties.

30 31

32

1. INTRODUCTION

33

The M4S6 highly symmetric sulfides of group 15 elements of the periodic table (M: P, As, Sb,

34

Bi) are examples of prototype cage like inorganic molecules. These serve as structural building

35

blocks of network like condensed matter and could be used to describe and predict the

36

physicochemical properties of various solid- and liquid-state systems [1]. Investigating the

37

properties of these molecules could reveal reaction mechanisms for the development of less

38

expensive catalysts as well as the formation of novel atomic-level controlled nanomaterials [2].

39

The description of the chemical bonding in the M4S6 molecules is a rather challenging task. The

40

knowledge of the structural parameters (bond lengths and bond angles) is inadequate to indicate

41

the strength of the metal-sulfur bond. It is essential to ascertain additional molecular properties

42

(especially in the case of more ionic systems), such as the population analysis, calculation of

43

the electrostatic potential surface and bond stretching frequencies [3–7]. From the

44

computational aspect, our conclusions are important in selecting the most accurate ‘low-cost’

45

level of theory in predicting the geometrical parameters and frontier orbitals, description of the

46

metal-sulfur bond, electrostatic potential surfaces and vibrational spectrum of metal sulfides.

47

The molecules M4S6 were considered as spherical top species with Td symmetry. To our

48

knowledge neither experimental structural data from electron diffraction studies nor vibrational

49

spectra of the gas phase Td sulfides have been published yet, so the existence of these species

50

is currently under consideration. Sulfides of group 15 of lower symmetry than Td have been

51

reported in the literature. Jason has suggested the possible structures of α-P4S6 (C1 symmetry),

52

β-P4S6 and γ-P4S6 (Cs symmetry) employing the

53

structurally characterized by Blachnik et al. [9] Furthermore, most of the arsenic sulfide

54

minerals are composed of cage-like molecules interacting through weak van der Waals (vdW)

55

forces. The uzonite mineral contains the cage molecules As4S5 [10]. The cyclo- anion Sb4S62−

56

has been obtained as its PPh4+ salt, where two exocyclic Sb-S bonds and an Sb-Sb bond are

2

P-NMR technique [8]. β-P4S6 has been

31

57

present [11]. In individual Sb2S3 nanowires embedded in anodic alumina templates

58

piezoelectric and ferroelectric properties have been observed [12]. Also, Sb2S3 crystals of the

59

urchin-like nanostructure, 3−4 μm in length and 30−150 nm in diameter have been obtained at

60

a mild reaction temperature [13]. Bi2S3 has a very anisotropic one-dimensional orthorhombic

61

structure considered as aligned (Bi4S6)n ribbons bound to each other by the vdW forces. It has

62

multiple functions in solar cells, such as a sensitizer, light absorber or electron acceptor material

63

[14]. In addition, the synthesis and the promising thermoelectric properties of highly oriented

64

bulk crystalline ingots of n-type bulk Bi2S3 doped with BiCl3 have been reported [15].

65

In this work, the formation of the M-S bond of group 15 sulfides M4S6 was characterized by

66

analyzing the electron density of the metal’s p valence orbitals and the 3p orbitals of sulfur

67

rearrangement between the two nuclei. In the literature has been reported charge transfer from

68

the 3p orbitals of sulfur to the p valence orbitals of M for sulfur complexes with M=Au, Ag, Cu

69

[4]. It is well known that phosphorous like nitrogen forms mostly covalent bonds while

70

compounds of arsenic, antimony, and bismuth are characterized by more ionic bonding.

71

Moreover, in molecules with M-S bonds, such as [PbII(S2COEt)n]2-n (n = 1,2,3,4), Ghosh et al.

72

[16] concluded that the PbII-S bond is formed by the 6p orbitals of PbII and the 3p of the S

73

atoms. Both PbII and MIII have the electronic structure ns2p0 but the s orbital of M has limited

74

contribution to the bond. The electron lone pair of the M atom is localized in hybrid orbital

75

having a higher percentage of s character. Along the P4S6 to Bi4S6 series, the localization of the

76

lone pair is significantly increased in an orbital with a higher percentage of s character.

77

The adamantane structure of P4S6 has been proposed by Gimarc and Ott but without any

78

experimental evidence [17]. Theoretical studies have been conducted on the Td form of the

79

phosphorus decasulfide P4S10 at HF/6-31G*, MP2/6-31G* and B3LYP/6-31G* levels of theory

80

[18]. The first reported calculations for As4S6 at Td symmetry were carried out by Fukui et al.

81

They performed obsolete ASMO calculations for all valence electrons (INDO type) and

3

82

proposed that the As4S6 unit is stable and is a reasonable candidate for a structural unit in As2S3

83

glass [19]. Babić et al. have made computations for As4S6 (Td) using the outdated Vosko

84

exchange-correlation parameterization without nonlocal terms correction and an s, p orbital

85

basis set without d-type functions for arsenic atoms [20]. Their theoretical gas phase results

86

were compared with the experimental XPS spectrum of amorphous As2S3 solid. The geometry

87

and vibrations of As4S6 and Sb4S6 of C3ν symmetry have been calculated at HF/SBK level of

88

theory [21]. Calculations of Bi2S3 ribbon-like nanostructures at B3LYP/def-SV(P) and

89

PBE/def-SV(P) have been reported [22].

90

In the present research work, we systematically investigated the modification of the structural,

91

vibrational and electronic properties of the group 15 sulfides M4S6 at Td symmetry. Moreover,

92

we assessed the performance of DFT and MP2 methods by comparing to CCSD results. The

93

conclusions from this work are beneficial for further experimental and computational studies.

94

Experimentally, they may be useful in the structural investigation of the isolated gas phase

95

species, besides solving complex structures of liquid, crystalline or amorphous phases. From

96

the computational aspect, they are important in selecting the most accurate reasonable-cost level

97

of theory in predicting the geometrical parameters, vibrations, HOMO-LUMO gaps, and

98

electrostatic potential surfaces of the studied group 15 sulfides. We validated eight functionals

99

BP86, M06L, B3LYP, M06, Μ06-2Χ, CAMB3LYP, wB97XD, B2PLYP which span from

100

conventional GGA functionals to modern dispersion corrected hybrid-meta-GGA functionals,

101

and MP2 ab initio method. The comparison of geometrical parameters was made based on the

102

results from the CCSD calculations.

103 104 105

4

106

2. MATERIALS AND METHODS

107 108

2.1. Programs

109

All calculations were performed with the Gaussian 09 (version C.01) program package [23], in

110

the Linux Opensuse Leap 42.3 environment. We initially considered the M4O6 (M = P, As, Sb

111

and Bi) optimized structures of Td symmetry which have been previously reported [3].

112

Employing the chemical editor Avogadro (version 1.1.1) [24] we replaced each oxygen atom

113

with sulfur in order to obtain the four molecules P4S6, As4S6, Sb4S6 and Bi4S6. Then, we

114

optimized all structures employing the molecular mechanics UFF force field [25] implemented

115

in Avogadro. These input geometries have been fully optimized at each level of theory (MP2,

116

CCSD and DFT), setting very tight optimization criteria (see supporting information where

117

listed the final optimized geometries). For all molecules, positive frequencies were obtained

118

corresponding to stable conformations at energy minima.

119

Analysis of the electrostatic potential surface has been performed by the program Multiwfn

120

3.4.1 [26]. All related parameters were calculated according to the equations from Murray et al.

121

[27]. The natural bond orbital population analysis was carried out with the NBO 3.1 program

122

[28] on wavefunctions calculated at the CCSD/LANL08(d) level of theory. The NBO method

123

considers localized bonds and lone electron pairs as the basic units of the molecular structure.

124

The program Multiwfn was also used to analyze NBO and NPA results. Harmonic vibrations

125

were assigned with the aid of Chemcraft and VEDA 4 programs [29,30]. The results were

126

visualized with Avogadro, VMD (Visual Molecular Dynamics) [31] and Chemcraft.

127 128

2.2. Methods

129

The following methods were utilized: (i) CCSD coupled cluster method [32,33] and (ii) MP2

130

based on the Møller-Plesset second-order perturbation theory [34]. (iii) BP86 is classified as a

131

generalized gradient approximation (GGA) and consists of the Becke88 exchange functional

5

132

[35] and the correlation functional Perdew86 [36]. (iv) M06L [37] is classified as a meta-

133

generalized gradient-approximation (meta-GGA) where the term ‘meta’ denotes dependence

134

on kinetic energy density. (v) B3LYP is a hybrid functional of generalized gradient

135

approximation (hybrid GGA). It consists of the Becke88 exchange functional [35] and the Lee-

136

Yang-Parr (LYP) correlation functional [38,39]. (vi) M06 and (vii) M06-2X, are hybrid

137

functionals of meta-generalized gradient approximation (hybrid meta-GGA) [40]. The M06-2X

138

method has been configured to include a medium-range correction [37]. (viii) CAMB3LYP and

139

(ix) wB97XD are hybrid exchange-correlation functionals with short and long-range correction.

140

CAMB3LYP consists of 19% Hartree-Fock (HF) and 81% Becke 1988 (B88) for the

141

interactions of short-range exchange, while for long range exchange, it consists of 65% Hartree-

142

Fock (HF) and 35% Becke 1988 (B88) [41]. The wB97XD consists of 100% exact long-range

143

exchange, 22% exact short-range exchange, one modified B97 density exchange functional for

144

short-range interactions, the B97 correlation density functional and empirical dispersion

145

corrections [42,43]. (x) B2PLYP is a double hybrid exchange-correlation functional. The first

146

part consists of Becke (B) exchange terms and Lee-Yang-Parr (LYP) correlation terms [38,39],

147

while the second one consists of exchange terms from Hartree-Fock (HF) and correlation terms

148

from the second order perturbation theory (PT2) [44].

149 150

2.3. Basis sets

151

The used basis sets for the sulfur, phosphorous, arsenic and heavier elements (Sb and Bi) are

152

the Los Alamos National Laboratory LANL2DZ [45–47] and its completely uncontracted basis

153

denoted as LANL08 [48]. The LANL08 basis sets for main group elements have been derived

154

from the Hay-Wadt LANL2DZ sets and correspond to triple-ζ valence orbital quality. The

155

choice of the basis sets is based on its ability to offer an effective core potential (ECP), which

156

reduces the number of electrons that are considered explicitly and speeds up the calculations. It

6

157

employs a core including for phosphorous and sulfur 10 electrons ([Ne]), for arsenic 28

158

electrons ([Ar] + 3d), for antimony 46 electrons ([Kr] + 4d) and for bismuth 78 electrons ([Xe]

159

+ 5d4f). For a more accurate description of the nature of the chemical bond between metal and

160

sulfur and for a better prediction of the vibrational energies, the addition of diffuse p-

161

function(s) as well as of polarization d- function(s) is necessary. A set of polarization functions

162

for the main group atoms has been determined by Gilbert and co-workers and these are denoted

163

as LANL08(d) and LANL2DZpd [49]. An all-electron, fully optimized contracted Gaussian-

164

basis set of triple zeta valence quality, named TZVp [50] for atoms P, S and As has been also

165

used for calculations on P4S6 and As4S6 molecules. All basis sets were obtained from the EMSL

166

Basis Set Library and the Basis Set Exchange (BSE) software [51,52].

167 168

3. RESULTS AND DISCUSSION

169

3.1. Molecular structure

170

The equilibrium structure of the four cage-like molecules M4S6, computed at

171

CCSD/LANL08(d) level, is depicted in Figure 1. The structural details obtained at various

172

methods are reported in Table 1. In the M4S6 (M= P, As, Sb, Bi) molecules at Td symmetry,

173

there are two groups of atoms equivalent by symmetry, the four metal and the six sulfur atoms.

174

There are one type of M-S bond and two bond angles (∠M-S-M, ∠S-M-S).

175

An increase in the length of the M-S bond across the 15th group was observed at all levels of

176

theory. For the CCSD/LANL08(d) optimized geometries, r(P-S) (2.158Å) < r(As-S) (2.274Å)

177

< r(Sb-S) (2.461Å) < r(Bi-S) (2.532Å). This increase is due to the different degree of covalent

178

bonding as well as to the fact that the heavier Sb and Bi atoms occupy a larger atomic volume.

179

The P-S bond length of β-P4S6 has a value equal to 2.145Å [9] and this is in excellent agreement

180

with the calculated value equal to 2.158Å at CCSD/LANL08(d) level of theory. The ∠M-S-M

181

is increased from M = P to M = Sb, whereas in Sb4S6 and Bi4S6 these angles are approximately

7

182

equal. The ∠S-M-S for P4S6 and As4S6 are approximately equal while they are decreasing for

183

Sb4S6 and Bi4S6.

184

The performance of popular density functionals has been investigated, following the

185

methodology presented in the DFT evaluation study of Minenkov et al. for the bond lengths of

186

ruthenium catalysts [53]. For each level of theory, the mean absolute error (𝜀𝑚.𝑎.𝑒. ) of the

187

structural parameters of the four molecules and the absolute error (𝜀𝑎.𝑒. ) of each molecule were

188

calculated by the following equations: 1

189

𝜀𝑚.𝑎.𝑒 = 𝑁 ∑𝑁 𝑖=1|𝑥𝑖 − 𝑐𝑖 |

(1)

190

𝜀𝑎.𝑒 = |𝑥𝑖 − 𝑐𝑖 |

(2)

191

where 𝑥𝑖 denotes the calculated bond length with DFT or MP2 methods and 𝑐𝑖 expresses the

192

calculated bond length values with the CCSD method.

193

The mean absolute error of bond lengths for the four M4S6 molecules relative to

194

CCSD/LANL08(d) is presented in Figure 2.

195

For all four molecules, it was found that MP2 method underestimates the bond lengths

196

(𝜀𝑚.𝑎.𝑒. = 0.008Å) relative to the CCSD. Similar bond elongation from MP2 to CCSD has been

197

reported on the sulfur-hydrogen bond (rSH) of H2S and the sulfur-sulfur bond in H2S dimers, as

198

well as an excellent agreement of CCSD calculated data with the experimental rSH bond length

199

within 0.001 Å [54]. Moreover, the CCSD method gives the best results for the geometries of

200

larger molecules such as SeO, SeCl, and AsO [55]. Helgaker et al. calculated MP2 geometries

201

with longer bonds than the CCSD ones for 19 molecules consisted of the light-weight elements

202

H, F, O, N and C [56] but this is not applicable to the studied molecules which consisted of

203

heavier atoms P, As, Sb, Bi, and S. Considering the above remarks we selected CCSD as the

204

reference method for our DFT calculations.

8

205

The most accurate functionals are the M06L, B2PLYP, CAMB3LYP and M06 with similar

206

performance according to the 𝜀𝑚.𝑎.𝑒. (Table S2). The ranking of the methods relative to CCSD

207

and based on the 𝜀𝑚.𝑎.𝑒. for all molecules is the following:

208

BP86< B3LYP< wB97XD< M06-2X< MP2< M06L≤ B2PLYP≤ CAMB3LYP≤ M06< CCSD.

209

In terms of 𝜀𝑎.𝑒 (see Table S2) the most accurate methods are wB97XD and CAMB3LYP for

210

P4S6, M06L and M06 for As4S6, B2PLYP and M06 for Sb4S6, and M06 and M06L for Bi4S6

211

(Figures S1-S4). The M06 and M06L methods show high accuracy for the molecules with the

212

higher-weight elements such as As, Sb and Bi while the wB97XD shows an excellent

213

performance for the lowest-weight P4S6.

214 215

3.2. Atomic charges

216

The natural atomic charges as calculated with natural population analysis (NPA) are presented

217

in Table 2. The M and S atoms are positively and negatively charged for all M4S6 molecules,

218

respectively. The charge increases from P to Bi, indicating an increase in the ionic character of

219

the bonds M-S and that the electrons’ density moves from metal to sulfur, as one could predict

220

on the basis of electronegativities. In particular, NPA shows the ability of sulfur atoms to host

221

part of the negative charge provided by the M atoms resulting in the stabilization of the

222

adamantane structure. Both MP2 and CCSD methods provide almost equal values for the

223

natural atomic charges as recorded in Table 2. The metal’s charge values are in the range 0.42

224

– 1.17 for M4S6 species showing a less ionic character of M-S than M-O bond for heavier

225

metals, compared with the values 0.35 - 2.05 from phosphorous to bismuth, for M4O6 species

226

[3].

227 228

3.3. Electrostatic Potential Analysis

229

The electrostatic potential (ESP) on the vdW molecular surface provides information about the

230

strength and the orientation of intermolecular interactions such as hydrogen or halogen bonding

9

231

[57,58] as well as the electrophilic and nucleophilic positions of the molecule where chemical

232

reactions are expected. The following parameters have been described in published work by

233

Murray et al. [27]. 𝑉̅𝑆+ and 𝑉̅𝑆− indicate the mean positive and negative ESP values on the vdW

234

surface respectively, Π is the mean deviation on the surface, which is an index of the internal

235

2 charge separation, and 𝜎𝑡𝑜𝑡 is the total ESP variance which is the sum of the positive 𝜎+2 and

236

the negative 𝜎−2 parts. The greater the 𝜎+2 and 𝜎−2 , the greater the molecule's tendency to interact

237

with other molecules through the positive and negative ESP domains. The degree of charge

238

balance equals ν and when the 𝜎+2 and 𝜎−2 are equal, then ν is maximized and equals 0.250. As

239

long as ν is closer to 0.250, the more likely it is that the molecule interacts with other ones

240

2 through the positive and negative regions to a similar extent. The product 𝜎𝑡𝑜𝑡 𝜈 is also a very

241

useful quantity; a large value indicates a relatively strong tendency to interact with other

242

molecules of the same kind electrostatically.

243

According to the data in Table 3, there is an increase in the values of all parameters V, d, Π,

244

2 2 𝜎𝑡𝑜𝑡 , 𝜎+2 , 𝜎−2 , 𝜈 and 𝜈𝜎𝑡𝑜𝑡 along the group 15, from M = P to M = Bi. The internal charge

245

separation is increased from P4S6 (𝛱 = 6.12 𝑘𝑐𝑎𝑙 𝑚𝑜𝑙 −1 ) to the heavier Bi4S6 (𝛱 =

246

16.14 𝑘𝑐𝑎𝑙 𝑚𝑜𝑙 −1 ). This separation is also confirmed by the charge values (Table 2) where

247

the charge difference between Bi and S atoms in Bi4S6 is greater than that between P and S

248

atoms in P4S6. The 𝜈 parameter shows the degree of charge balance and if it receives the

249

maximum value 𝜈 = 0.25 then the molecule interacts with other molecules to the same extent

250

with its positive and its negatively charged region. Also, the 𝜈 parameter increases from P4S6

251

to Bi4S6, so more intermolecular interactions with different molecules for the heavier Bi4S6 are

252

2 expected. Moreover, a high value of 𝜈𝜎𝑡𝑜𝑡 indicates that a molecule has strong electrostatic

253

2 interactions among other molecules of the same species. The 𝜈𝜎𝑡𝑜𝑡 increases from P4S6

254

2 2 (𝜈𝜎𝑡𝑜𝑡 = 3.55 (𝑘𝑐𝑎𝑙 𝑚𝑜𝑙 −1 )2) to Bi4S6 (𝜈𝜎𝑡𝑜𝑡 = 19.04 (𝑘𝑐𝑎𝑙 𝑚𝑜𝑙 −1 )2) whereby Bi4S6

255

molecules interact strongly with each other. For example, it has been found that in saturated

10

256

2 hydrocarbons 𝜈𝜎𝑡𝑜𝑡

is about 1 (𝑘𝑐𝑎𝑙 𝑚𝑜𝑙 −1 )2

2 while for formamide is 𝜈𝜎𝑡𝑜𝑡 =

257

62.5 (𝑘𝑐𝑎𝑙 𝑚𝑜𝑙 −1 )2 [27] so that it will have strong electrostatic forces through its positive and

258

negative regions and formamide molecules should interact strongly with each other. That was

259

confirmed experimentally after the investigation of the N-H stretching domains at the infrared

260

and Raman spectra in the solid and liquid state of formamide [59]. The Figures S5-S8 illustrate

261

the electrostatic potential surfaces of M4S6 molecules at CCSD/LANL08(d) level of theory. The

262

colors are related to the electrostatic potential values. Red color indicates electronic deficient

263

areas (𝑉(𝑟) > 0) and blue indicates electron rich areas (𝑉(𝑟) < 0).

264 265

3.4. Bonding analysis

266

The analysis of the chemical bonding and of the HOMO-LUMO orbitals was based on results

267

from the CCSD/LANL08(d) method. For the compounds M4S6, the simple σ bond between M

268

and S can be written as: 𝜎𝑀𝑆 = 𝑐𝑀 ℎ𝑀 + 𝑐𝑆 ℎ𝑆

269

(3)

270

2 where 𝑐𝑀 and 𝑐𝑆 are the polarity coefficients of the hybrid orbitals ℎ𝑀 and ℎ𝑆 , while 𝑐𝑀 + 𝑐𝑆2 =

271

1. The bond ionicity parameter provided by the Eq. (4) and quantifies the polarity of the bond

272

[60].

273

𝑐 2 −𝑐 2

𝑆 𝑖𝑀𝑆 = 𝑐𝑀 2 +𝑐 2 𝑀

(4)

𝑆

274

For the P4S6 molecule, all 𝜎𝑃𝑆 bonds are equivalent and each one is expressed according to Eq.

275

(3) as 𝜎𝑀𝑆 = 0.635(𝑠𝑝6.93 𝑑0.14 )𝑃 + 0.772(𝑠𝑝6.00 𝑑 0.07 )𝑆 . Also, each P atom has a lone pair of

276

electrons in an orbital with hybridization 𝑠𝑝0.59 (62.83% 3s of P) while each S atom has two

277

lone pairs of electrons, the first one in 𝑠𝑝0.39 (71.65% 3s of S) and the second one in the 3p of

278

S (99.86%). The |iMS | parameter was found equal to 0.193. In each of the other molecules

279

As4S6, Sb4S6 and Bi4S6, the σ bonds are equivalent, and the lone electron pairs are in a similar

280

configuration to P4S6, i.e. one in M and two in S. The results are summarized in Table 4.

11

281

Based on the NBO theory, it follows from the values of the bond ionicity |iMS | that the most

282

covalent bond M-S exists in P4S6 with |iMS | = 0.193. As M varies from P to Bi, the bond

283

becomes more polar, since for M = Bi, |iMS | = 0.464. For all the studied species it was found

284

that 𝑐𝑀 < 𝑐𝑆 and the bond is more polarized to the S atoms. In addition, the M-S bond is formed

285

from the p valence orbitals of M and the 3p orbital of S. According to Table 4, the participation

286

of the p orbitals of M and of S to the 𝜎𝑀𝑆 bond is increased from P4S6 (𝜎𝑀𝑆 =

287

0.635(𝑠𝑝6.93 𝑑 0.14 )𝑃 + 0.772(𝑠𝑝6.00 𝑑0.07 )𝑆 )

288

0.856(𝑠𝑝7.58 𝑑0.04 )𝑆 ).

289

In molecules with M-S bonds such as [PbII(S2COEt)n]2-n (n=1,2,3,4), Ghosh et al. concluded

290

that the PbII-S bond is formed from the 6p orbitals of PbII and the 3p ones of S atoms [16]. A

291

similar conclusion about the M-S bond was reached for the studied M4S6 molecules in this

292

work. The NBO analysis of the 𝜎𝑀𝑆 bond showed that the donor NBOs are composed mainly

293

of the 3p orbital of sulfur while the acceptor NBOs are mostly comprised of the valence p

294

orbitals of M. Therefore, a general tendency is observed in the M-S bond of association between

295

p valence orbital of the metal and the 3p of sulfur.

296

The lone electron pair of M denoted as nM is found in hybrid orbitals with higher s character.

297

Along the group 15 of the periodic table, the localization of the lone pair increases significantly

298

in the s orbital of M, since in the P4S6 it is 62.83% s while for the Bi4S6 it is 86.59% s. Similarly,

299

the first lone electron pair of S denoted as nS , is in a hybrid orbital with more s character.

300

From M = P to M = Bi there is small increase in s character: for P4S6 is 71.65% while for Bi4S6

301

it is 76.74%. The second lone electron pair of S, nS

302

molecules.

to

Bi4S6

(𝜎𝐵𝑖𝑆 = 0.518(𝑠𝑝21.00 𝑑0.07 )𝐵𝑖 +

(𝜎)

(𝜎)

(𝜋)

303

12

is located on the 3p orbital in all M4S6

304

3.5. Frontier orbitals analysis

305

The analysis of the frontier molecular orbitals calculated at the CCSD/LANL08(d) level of

306

theory for all M4S6 molecules follows. In all studied molecules, there are three energy-

307

degenerated HOMO orbitals having t1 symmetry (indicated as MO-26, MO-27 and MO-28) and

308

two degenerated LUMO orbitals of e symmetry (indicated as MO-29 and MO-30). These five

309

MO orbitals are presented in Figures 3 and S9-S11.

310

The participation of the atomic orbitals to the aforementioned five frontier orbitals has been

311

analyzed.

312

P4S6: The 3py, 3px and 3pz atomic orbitals of P atoms have the largest contribution (by 34.32%)

313

to the three HOMO orbitals of P4S6 MO-26, MO-27 and MO-28 respectively. Also, these three

314

HOMO orbitals consisted of the 3s orbital of the P atoms, with a contribution of 10.71% and of

315

the 3s of the S atoms with 1.46% participation. The two LUMO orbitals do not consist of 3s of

316

P but consisted by 12.72% of 3s of S. Also, MO-29 is 39.82% of 3py and 10.90% of 3pz of P

317

whereas MO-30 is 30.80% 3px and 28.94% 3pz of P.

318

As4S6: For the As4S6, the 4s of the As atoms contributed a total of 11.97% and the 3s of the S

319

atoms a total of 1.08%. The 4px, 4pz and 4py of As were significantly involved in the three

320

HOMO orbitals MO-26, MO-27 and MO-28 by 30.80% respectively. The two LUMO orbitals

321

did not consist of 4s of As but 11.61% of 3s of S. Also, MO-29 consisted of 37.53% of 4pz and

322

26.26% of 4py of As while MO-30 from 42.20% 4px and 16.94% 4py of As.

323

Sb4S6: In Sb4S6 the three HOMO orbitals consisted mainly of 5py, 5pz and 5px of Sb which

324

contribute 27.69% to MO-26, MO-27 and MO-28 respectively. Also, the three HOMOs

325

consisted of the 5s of the Sb by 16.58%. The two LUMOs were not composed of 5s of Sb but

326

of 9.83% of the 3s of S. The MO-29 consisted of 41.45% of 5py of Sb and 29.86% of 5pz of Sb

327

while MO- 30 from 47.22% 5px and 18.30% 5pz of Sb.

13

328

Bi4S6: For the Bi4S6 the three HOMO orbitals consisted of the 6s orbitals of Bi by 12.04% and

329

of the 3s of S atoms by 0.52%. The 6py, 6pz and 6px participated in MO-26, MO-27 and MO-

330

28 by 20.38% respectively. The two LUMOs were not composed of 6s of Bi but of 9.10% of

331

3s of S. Furthermore, MO-29 consisted of 48.00% of 6px and 14.61% of 6pz of Bi. The MO-30

332

was composed of 38.54% 6py and 33.57% 6pz of Bi atoms.

333 334

All contributions of the atomic orbitals to the aforementioned HOMOs and LUMOs for M4S6

335

molecules are provided in the Table 5 of this paper. In all studied molecules, it was found that

336

their HOMO consisted mainly of the p valence orbitals of the M atoms and of the 3p orbitals of

337

the S atoms. Along the group 15, the contribution of 3p of S atoms is significantly increased,

338

the 3py of the P4S6 participates 37.64% in MO-26 while in Bi4S6 3py participates by 55.41%. In

339

addition, the contribution of the p valence orbitals of M to the LUMOs is significant in all M4S6

340

molecules.

341 342

For the energy gaps between HOMO and LUMO orbitals at the CCSD/LANL08(d) level of

343

theory, it was found that from M = P to M = As there is a decrease in these values whereas there

344

is no significant difference between Sb4S6 and Bi4S6 molecules. That is, the following

345

relationship applies to the HOMO-LUMO energy gaps:

346 347

𝛥𝐸𝑃4 𝑆6 (10.5 𝑒𝑉) > 𝛥𝐸𝐴𝑠4 𝑆6 (10.0 𝑒𝑉) > 𝛥𝐸𝑆𝑏4 𝑆6 , 𝛥𝐸𝐵𝑖4 𝑆6 (9.3 𝑒𝑉)

348 349

Since all 𝛥𝐸𝑀4 𝑆6 values are high, M4S6 molecules are expected to be chemically stable.

350

14

351

3.6. Molecular Polarizability

352

The perturbation of electron density when a molecule interacts with the electric field of

353

radiation, ions, polar molecules, etc. is fundamental to understand the behavior of molecules in

354

chemical reactions, their solvation properties, the recognition processes and spectroscopic

355

properties. Electric polarizability is currently of importance as it is extensively used to model

356

intermolecular interactions [61], basic molecular characteristics as the acidity and basicity [62],

357

hardness and softness [63–65] and chemical reactivity.

358

Our polarizability calculations were performed at the CCSD(full)/LANL08(d) optimized

359

geometry, using the MP2(full)/LANL08(d) theoretical level. Due to high symmetry there is

360

only one independent component of the polarizability tensor: αxx = αyy = αzz = 𝛼̅.

361

In atomic units the calculated values of the mean polarizability are:

362

𝛼̅(𝑃4 𝑆6 ) = 193.91, 𝛼̅(𝐴𝑠4 𝑆6 ) = 217.87, 𝛼̅(𝑆𝑏4 𝑆6 ) = 267.46, 𝛼̅(𝐵𝑖4 𝑆6 ) = 285.82 e2α02Eh-1.

363

Similar calculations were performed for the M4O6 molecules:

364 365

𝛼̅(𝑃4 𝑂6 ) = 89.106, 𝛼̅(𝐴𝑠4 𝑂6 ) = 107.17, 𝛼̅(𝑆𝑏4 𝑂6 ) = 142.24, 𝛼̅(𝐵𝑖4 𝑂6 ) = 155.27 e2α02Eh-1

366

A quantity related to polarizability is the polarizability volume, α', defined as α' = α/(4πε0)

367

For M4S6 molecules these volumes have been calculated:

368

α'(P4S6)= 28.705, α'(As4S6)= 32.252, α'(Sb4S6)= 39.593, α'(Bi4S6)= 42.312 Å3

369

As it is expected polarizability and polarizability volume increase as the size of atoms (P to Bi

370

and O to S) increases.

371 372

3.7. Vibrational properties

373

The highly symmetric M4S6 Td molecules have the following irreducible representation of the

374

fundamental vibrations:

375

Γvib. = 2Α1(R) + 2Ε(R) + 2T1(i.a.) + 4T2(IR,R)

15

376

The vibrational modes having A1, E and T2 symmetry are Raman (R) active and only the ones

377

having T2 symmetry are IR active (IR). The vibrational modes having T1 symmetry are both IR

378

and Raman forbidden (i.a.).

379

In Table 6 the harmonic wavenumbers calculated at the optimized geometry and their

380

symmetries, are presented.

381

A non-linear molecule with N atoms exhibits 3N - 6 normal modes of vibration. There are 24

382

fundamental vibrations - symmetric and antisymmetric – for all M4S6 cages. Each fundamental

383

mode can be expressed by a normal coordinate. Then, those coordinates are transformed to

384

internal ones, which are a superposition of multiple local modes including stretching, bending

385

and torsion [29,30]. The Greek letters ν, β, δ, τ denote stretching, in plane bending, out of plane

386

bending and torsion modes respectively (Table 6). By using the symmetry considerations and

387

theoretical results for IR and Raman intensities, it is possible to assign the predicted energies

388

of vibrations to specific vibrational modes.

389

For all systems investigated here and for all calculated modes, the corresponding vibrational

390

wavenumbers decrease when the atomic number of the central atom increases. For the M-S

391

stretching vibrations this decrease is expected since the quadratic M-S force constants decrease

392

in the series from P to Bi whereas the reduced mass increases reinforcing the behavior expected

393

from the quadratic force constant.

394

There is also a correlation between bond angles and hybridization. The percentage of s character

395

of the M to the 𝜎𝑀𝑆 bond decreases from P to Bi. However, the contribution of the p orbitals of

396

M increases from P to Bi. In addition, we found that the angle ∠S-M-S decreases from P4S6 to

397

Bi4S6. Therefore, P4S6 with the highest %s contribution of phosphorus to the 𝜎𝑀𝑆 bond, has the

398

largest value of ∠S-P-S equal to 106.97°. This trend is in agreement with the prediction of Bent

399

rule, which was originally derived from sp-hybridized main group elements [66].

16

400

4. CONCLUSIONS

401 402

In the present work, the structural, electronic and vibrational properties of the M4S6, M = P, As,

403

Sb, Bi inorganic clusters were investigated by means of computer simulation methods. The

404

calculations were performed with the MP2, CCSD and DFT methods employing the basis set

405

LanL08(d) with ECP for the P, As, Sb, Bi, S atoms. Comparing with the available CCSD

406

theoretical data, we concluded that M06 and M06L functionals show high accuracy for

407

molecules with the higher-weight elements such as As, Sb, and Bi while the wB97XD and

408

CAMB3LYP show an excellent performance for the lowest-weight P4S6 concerning the

409

prediction of the structural data. We also present for the first time data for the hypothetical

410

Sb4S6 and Bi4S6 molecules. The vibrational spectral analysis is presented for the whole series

411

of group 15 metal sulfides. A structural ‘relaxation’ ongoing from phosphorous to bismuth

412

indicates an increase of ionic character and in some extent can explain the instability of these

413

compounds when the atomic number of the metal increases.

17

414

References

415

[1]

416 417

doi:10.1126/science.196.4292.839. [2]

418 419

E.L. Muetterties, Molecular Metal Clusters, Science (80-. ). 196 (1977) 839–848.

Z. Luo, A.W. Castleman, S.N. Khanna, Reactivity of Metal Clusters, Chem. Rev. 116 (2016) 14456– 14492. doi:10.1021/acs.chemrev.6b00230.

[3]

A. Chrissanthopoulos, C. Pouchan, A density functional investigation of the structural and vibrational

420

properties of the highly symmetric molecules M4O6, M4O10 (M=P, As, Sb, Bi), Vib. Spectrosc. 48 (2008)

421

135–141. doi:10.1016/j.vibspec.2008.02.012.

422

[4]

423 424

clusters, J. Phys. Chem. A. 114 (2010) 9212–9221. doi:10.1021/jp100423b. [5]

425 426

A.H. Pakiari, Z. Jamshidi, Nature and strength of MS bonds (M = Au, Ag, and Cu) in binary alloy gold

A.H. Pakiari, Z. Jamshidi, Interaction of Coinage Metal Clusters with Chalcogen Dihydrides, J. Phys. Chem. A. 112 (2008) 7969–7975. doi:10.1021/jp804033w.

[6]

E.S. Yandanova, D.M. Ivanov, M.L. Kuznetsov, A.G. Starikov, G.L. Starova, V.Y. Kukushkin,

427

Recognition of S···Cl Chalcogen Bonding in Metal-Bound Alkylthiocyanates, Cryst. Growth Des. 16

428

(2016) 2979–2987. doi:10.1021/acs.cgd.6b00346.

429

[7]

B.M. Barry, B.W. Stein, C.A. Larsen, M.N. Wirtz, W.E. Geiger, R. Waterman, R.A. Kemp, Metal

430

Complexes (M = Zn, Sn, and Pb) of 2-Phosphinobenzenethiolates: Insights into Ligand Folding and

431

Hemilability, Inorg. Chem. 52 (2013) 9875–9884. doi:10.1021/ic400990n.

432

[8]

M. Jason, Transfer of Sulfur from Arsenic and Antimony Sulfides to Phosphorus Sulfides. Rational

433

Syntheses of Several Less-Common P4Sn Species, Inorg. Chem. 36 (1997) 2641–2646.

434

doi:10.1021/ic9614881.

435

[9]

R. Blachnik, U. Peukert, A. Czediwoda, B. Engelen, K. Boldt, Die Molekulare Zusammensetzung von

436

erstarrten Phosphor-Schwefel-Schmelzen und die Kristallstruktur von β-P4S6, Zeitschrift Für Anorg. Und

437

Allg. Chemie. 621 (1995) 1637–1643. doi:10.1002/zaac.19956211004.

438

[10]

439 440

L. Bindi, V. Popova, P. Bonazzi, Uzonite, As4S5, From the type locality: single-crystal X-ray study and effects of exposure to light, Can. Mineral. 41 (2003) 1463–1468. doi:10.2113/gscanmin.41.6.1463.

[11]

T.M. Martin, G.L. Schimek, W.T. Pennington, J.W. Kolis, Synthesis of two new antimony sulfide

441

clusters: structures of [PPh4]2[Sb6S6] and [PPh4]2[Sb4S6], J. Chem. Soc. Dalt. Trans. (1995) 501–502.

442

doi:10.1039/DT9950000501.

443

[12]

J. Varghese, S. Barth, L. Keeney, R.W. Whatmore, J.D. Holmes, Nanoscale Ferroelectric and

444

Piezoelectric Properties of Sb2S3 Nanowire Arrays, Nano Lett. 12 (2012) 868–872.

445

doi:10.1021/nl2039106.

446

[13]

447 448

N. Maiti, S.H. Im, Y. Lee, S. Il Seok, Urchinlike Nanostructure of Single-Crystalline Nanorods of Sb2S3 Formed at Mild Reaction Condition, 2012. doi:10.1021/am301141q.

[14]

D. Han, M.-H. Du, C.-M. Dai, D. Sun, S. Chen, Influence of defects and dopants on the photovoltaic

449

performance of Bi2S3: first-principles insights, J. Mater. Chem. A. 5 (2017) 6200–6210.

450

doi:10.1039/C6TA10377D.

451

[15]

452 453

K. Biswas, L.-D. Zhao, M.G. Kanatzidis, Tellurium-Free Thermoelectric: The Anisotropic n-Type Semiconductor Bi2S3, Adv. Energy Mater. 2 (n.d.) 634–638. doi:10.1002/aenm.201100775.

[16]

D. Ghosh, K. Sen, S. Bagchi, A.K. Das, The nature of lead-sulfur interaction in [PbII(S2COEt)n]2-n (n =

18

454

1,2,3,4) complexes: topological exploration and formation analysis, New J. Chem. 37 (2013) 1408–1416.

455

doi:10.1039/C3NJ41105B.

456

[17]

B.M. Gimarc, J.J. Ott, Topologically Determined Charge Distributions and the Chemistry of Some Cage-

457

Type Structures Related to Adamantane, J. Am. Chem. Soc. 108 (1986) 4298–4303.

458

doi:10.1021/ja00275a010.

459

[18]

460 461

THEOCHEM. 487 (1999) 267–274. doi:10.1016/S0166-1280(99)00068-8. [19]

462 463

[20]

[21]

J.A. Tossell, Theoretical studies on As and Sb sulfide molecules, Miner. Spectrosc. A Tribut. to Roger G. Burn. Geochemical Soc. Spec. Publ. (1996) 9–19.

[22]

468 469

D. Babić, S. Rabii, J. Bernholc, Structural and electronic properties of arsenic chalcogenide molecules, Phys. Rev. B. 39 (1989) 10831–10838. doi:10.1103/PhysRevB.39.10831.

466 467

A. Tachibana, T. Yamabe, M. Miyake, K. Tanaka, H. Kato, K. Fukui, Electronic behavior of amorphous chalcogenide models, J. Phys. Chem. 82 (1978) 272–277. doi:10.1021/j100492a004.

464 465

J.O. Jensen, D. Zeroka, Theoretical studies of the infrared and Raman spectra of P 4S10, J. Mol. Struct.

V. Calzia, G. Malloci, G. Bongiovanni, A. Mattoni, Electronic properties and quantum confinement in Bi2S3 ribbon-like nanostructures, J. Phys. Chem. C. 117 (2013) 21923–21929. doi:10.1021/jp405740b.

[23]

M.J. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb, J.R. Cheeseman, G. Scalmani, V.

470

Barone, B. Mennucci, G.A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H.P. Hratchian, A.F. Izmaylov,

471

J. Bloino, G. Zheng, J.L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M.

472

Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J.A. Montgomery, Jr., J.E. Peralta, F.

473

Ogliaro, M. Bearpark, J.J. Heyd, E. Brothers, K.N. Kudin, V.N. Staroverov, T. Keith, R. Kobayashi, J.

474

Normand, K. Raghavachari, A. Rendell, J.C. Burant, S.S. Iyengar, J. Tomasi, M. Cossi, N. Rega, J.M.

475

Millam, M. Klene, J.E. Knox, J.B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R.E.

476

Stratmann, O. Yazyev, A.J. Austin, R. Cammi, C. Pomelli, J.W. Ochterski, R.L. Martin, K. Morokuma,

477

V.G. Zakrzewski, G.A. Voth, P. Salvador, J.J. Dannenberg, S. Dapprich, A.D. Daniels, O. Farkas, J.B.

478

Foresman, J.V. Ortiz, J. Cioslowski, D.J. Fox, Gaussian 09, Revision C.01, Wallingford CT, (2010).

479

[24]

M.D. Hanwell, D.E. Curtis, D.C. Lonie, T. Vandermeerschd, E. Zurek, G.R. Hutchison, Avogadro: An

480

advanced semantic chemical editor, visualization, and analysis platform, J. Cheminform. 4 (2012) 1–17.

481

doi:10.1186/1758-2946-4-17.

482

[25]

A.K. Rappe, C.J. Casewit, K.S. Colwell, W.A. Goddard, W.M. Skiff, UFF, a full periodic table force

483

field for molecular mechanics and molecular dynamics simulations, J. Am. Chem. Soc. 114 (1992)

484

10024–10035. doi:10.1021/ja00051a040.

485

[26]

486 487

T. Lu, F. Chen, Multiwfn: A multifunctional wavefunction analyzer, J. Comput. Chem. 33 (2012) 580– 592. doi:10.1002/jcc.22885.

[27]

J.S. Murray, T. Brinck, P. Lane, K. Paulsen, P. Politzer, Statistically-based interaction indices derived

488

from molecular surface electrostatic potentials: a general interaction properties function (GIPF), J. Mol.

489

Struct. THEOCHEM. 307 (1994) 55–64. doi:10.1016/0166-1280(94)80117-7.

490

[28]

E.D. Glendening, A.E. Reed, J.E. Carpenter, F. Weinhold, NBO Version 3.1, n.d.

491

[29]

M.H. Jamroz, Vibrational Energy Distribution Analysis VEDA 4, (2010).

492

[30]

M.H. Jamróz, Vibrational Energy Distribution Analysis (VEDA): Scopes and limitations, Spectrochim.

493

Acta Part A Mol. Biomol. Spectrosc. 114 (2013) 220–230. doi:10.1016/j.saa.2013.05.096.

19

494

[31]

495 496

W. Humphrey, A. Dalke, K. Schulten, VMD: visual molecular dynamics., J. Mol. Graph. 14 (1996) 2728,33-38. doi:10.1016/0263-7855(96)00018-5.

[32]

J. Paldus, J. Čížek, I. Shavitt, Correlation Problems in Atomic and Molecular Systems. IV. Extended

497

Coupled-Pair Many-Electron Theory and Its Application to the BH3 Molecule, Phys. Rev. A. 5 (1972)

498

50–67. doi:10.1103/PhysRevA.5.50.

499

[33]

J.A. Pople, M. Head‐Gordon, K. Raghavachari, Quadratic configuration interaction. A general technique

500

for determining electron correlation energies, J. Chem. Phys. 87 (1987) 5968–5975.

501

doi:10.1063/1.453520.

502

[34]

503 504

(1934) 618–622. doi:10.1103/PhysRev.46.618. [35]

505 506

A.D. Becke, Density-functional exchange-energy approximation with correct asymptotic behavior, Phys. Rev. A. 38 (1988) 3098–3100. doi:10.1103/PhysRevA.38.3098.

[36]

507 508

C. Møller, M.S. Plesset, Note on an approximation treatment for many-electron systems, Phys. Rev. 46

J.P. Perdew, Density-functional approximation for the correlation energy of the inhomogeneous electron gas, Phys. Rev. B. 33 (1986) 8822–8824. doi:10.1103/PhysRevB.33.8822.

[37]

Y. Zhao, D.G. Truhlar, A new local density functional for main-group thermochemistry, transition metal

509

bonding, thermochemical kinetics, and noncovalent interactions, J. Chem. Phys. 125 (2006) 194101.

510

doi:10.1063/1.2370993.

511

[38]

512 513

functionals, J. Chem. Phys. 107 (1997) 8554–8560. doi:10.1063/1.475007. [39]

514 515

A.D. Becke, Density-functional thermochemistry. V. Systematic optimization of exchange-correlation

C. Lee, W. Yang, R.G. Parr, Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density, Phys. Rev. B. 37 (1988) 785–789. doi:10.1103/PhysRevB.37.785.

[40]

Y. Zhao, D.G. Truhlar, The M06 suite of density functionals for main group thermochemistry,

516

thermochemical kinetics, noncovalent interactions, excited states, and transition elements: two new

517

functionals and systematic testing of four M06-class functionals and 12 other function, Theor. Chem.

518

Acc. 120 (2008) 215–241. doi:10.1007/s00214-007-0310-x.

519

[41]

T. Yanai, D.P. Tew, N.C. Handy, A new hybrid exchange–correlation functional using the Coulomb-

520

attenuating method (CAM-B3LYP), Chem. Phys. Lett. 393 (2004) 51–57.

521

doi:/10.1016/j.cplett.2004.06.011.

522

[42]

523 524

functionals, J. Chem. Phys. 128 (2008) 18315032. doi:10.1063/1.2834918. [43]

525 526

[44]

[45]

533

P.J. Hay, W.R. Wadt, Ab initio effective core potentials for molecular calculations. Potentials for K to Au including the outermost core orbitale, J. Chem. Phys. (1985). doi:10.1063/1.448975.

[46]

531 532

S. Grimme, Semiempirical hybrid density functional with perturbative second-order correlation, J. Chem. Phys. 124 (2006) 03408. doi:10.1063/1.2148954.

529 530

J.-D. Chai, M. Head-Gordon, Long-range corrected hybrid density functionals with damped atom-atom dispersion corrections., Phys. Chem. Chem. Phys. 10 (2008) 6615–6620. doi:10.1039/b810189b.

527 528

J. Da Chai, M. Head-Gordon, Systematic optimization of long-range corrected hybrid density

W.R. Wadt, P.J. Hay, Ab initio effective core potentials for molecular calculations. Potentials for main group elements Na to Bi, J. Chem. Phys. (1985). doi:10.1063/1.448800.

[47]

P.J. Hay, W.R. Wadt, Ab initio effective core potentials for molecular calculations. Potentials for the transition metal atoms Sc to Hg, J. Chem. Phys. (1985). doi:10.1063/1.448799.

20

534

[48]

535 536

L.E. Roy, P.J. Hay, R.L. Martin, Revised Basis Sets for the LANL Effective Core Potentials, J. Chem. Theory Comput. 4 (2008) 1029–1031. doi:10.1021/ct8000409.

[49]

C.E. Check, T.O. Faust, J.M. Bailey, B.J. Wright, T.M. Gilbert, L.S. Sunderlin, Addition of Polarization

537

and Diffuse Functions to the LANL2DZ Basis Set for P-Block Elements, J. Phys. Chem. A. 105 (2001)

538

8111–8116. doi:10.1021/jp011945l.

539

[50]

540 541

quality for atoms Li to Kr, J. Chem. Phys. 100 (1994) 5829–5835. doi:10.1063/1.467146. [51]

542 543

A. Schäfer, C. Huber, R. Ahlrichs, Fully optimized contracted Gaussian basis sets of triple zeta valence

D. Feller, The role of databases in support of computational chemistry calculations, J. Comput. Chem. 17 (1996) 1571–1586. doi:10.1002/(SICI)1096-987X(199610)17:133.0.CO;2-P.

[52]

K.L. Schuchardt, B.T. Didier, T. Elsethagen, L. Sun, V. Gurumoorthi, J. Chase, J. Li, T.L. Windus, Basis

544

Set Exchange:  A Community Database for Computational Sciences, J. Chem. Inf. Model. 47 (2007)

545

1045–1052. doi:10.1021/ci600510j.

546

[53]

Y. Minenkov, A. Singstad, G. Occhipinti, V.R. Jensen, The accuracy of DFT-optimized geometries of

547

functional transition metal compounds: a validation study of catalysts for olefin metathesis and other

548

reactions in the homogeneous phase, Dalt. Trans. 41 (2012) 5526–5541. doi:10.1039/C2DT12232D.

549

[54]

550 551

K.H. Lemke, Structure and binding energy of the H2S dimer at the CCSD(T) complete basis set limit, J. Chem. Phys. 146 (2017). doi:10.1063/1.4985094.

[55]

D.R. Urban, J. Wilcox, A theoretical study of properties and reactions involving arsenic and selenium

552

compounds present in coal combustion flue gases., J. Phys. Chem. A. 110 (2006) 5847–52.

553

doi:10.1021/jp055564+.

554

[56]

555 556

standard electronic wave functions, J. Chem. Phys. 106 (1997) 6430–6440. doi:10.1063/1.473634. [57]

557 558

[58]

[59]

[60]

[61]

A.D. Buckingham, Permanent and Induced Molecular Moments and Long-Range Intermolecular Forces, in: Adv. Chem. Phys., John Wiley & Sons, Ltd, 2007: pp. 107–142. doi:10.1002/9780470143582.ch2.

[62]

567 568

F. Weinhold, C. Landis, Valency And Bonding A Natural Bond Orbital Donor-Acceptor Perspective, Cambridge University Press, New York, 2005.

565 566

D.J. Gardiner, A.J. Lees, B.P. Straughan, A study of intermolecular hydrogen bonding in formamide by vibrational spectroscopy, J. Mol. Struct. 53 (1979) 15–24. doi:10.1016/0022-2860(79)80320-8.

563 564

T. Lu, S. Manzetti, Wavefunction and reactivity study of benzo[a]pyrene diol epoxide and its enantiomeric forms, Struct. Chem. 25 (2014) 1521–1533. doi:10.1007/s11224-014-0430-6.

561 562

T. Lu, F. Chen, Quantitative analysis of molecular surface based on improved Marching Tetrahedra algorithm, J. Mol. Graph. Model. 38 (2012) 314–323. doi:10.1016/j.jmgm.2012.07.004.

559 560

T. Helgaker, J. Gauss, P. Jo/rgensen, J. Olsen, The prediction of molecular equilibrium structures by the

A.D. Headley, Substituent effects on the basicity of dimethylamines, J. Am. Chem. Soc. 109 (1987) 2347–2348. doi:10.1021/ja00242a017.

[63]

M. Berkowitz, R.G. Parr, Molecular hardness and softness, local hardness and softness, hardness and

569

softness kernels, and relations among these quantities, J. Chem. Phys. 88 (1988) 2554–2557.

570

doi:10.1063/1.454034.

571

[64]

572 573

A. Vela, J.L. Gazquez, A relationship between the static dipole polarizability, the global softness, and the fukui function, J. Am. Chem. Soc. 112 (1990) 1490–1492. doi:10.1021/ja00160a029.

[65]

T.K. Ghanty, S.K. Ghosh, A Density Functional Approach to Hardness, Polarizability, and Valency of

21

574 575 576

Molecules in Chemical Reactions, J. Phys. Chem. 100 (1996) 12295–12298. doi:10.1021/jp960276m. [66]

H.A. Bent, An Appraisal of Valence-bond Structures and Hybridization in Compounds of the First-row elements., Chem. Rev. 61 (1961) 275–311. doi:10.1021/cr60211a005.

577

22

Figure 1. The adamantane structure of the M4S6 molecules at Td symmetry, as computed at CCSD/LANL08(d).

Figure 2. Mean absolute error for the four M4S6 molecules relative to CCSD/LANL08(d).

24

e

t1

Figure 3. Molecular orbital diagrams of HOMO and LUMO orbitals of P4S6.

25

Table 1. Optimized geometrical parameters of M4S6 molecules calculated at various levels of theory, using the LANL08(d) or TZVp [in brackets] basis sets. Structural Molecule

Parameter r(P-S) (Å)

P4S6

Bi4S6

2.149

CCSD

wB97XD

M06-2X

M06L

M06

BP86

CAMB3LYP

B2PLYP

B3LYP

2.158

2.156

2.154

2.169

2.168

2.187

2.156

2.164

2.177

∠P-S-P (°)

114.271 [113.999]

114.282

114.445

114.683

114.610

114.716

114.430

114.432

114.407

114.482

∠S-P-S (°)

106.969 [107.115]

106.963

106.874

106.744

106.784

106.726

106.882

106.881

106.894

106.854

[2.248]

2.274

2.265

2.265

2.279

2.278

2.301

2.268

2.279

2.291

∠As-S-As (°)

115.748 [115.301]

115.860

116.118

116.254

116.112

116.298

115.611

116.123

115.840

115.876

∠S-As-S (°)

106.159 [106.406]

106.097

105.953

105.877

105.956

105.853

106.234

105.950

106.108

106.088

2.454

2.461

2.444

2.450

2.468

2.463

2.489

2.454

2.467

2.477

∠Sb-S-Sb (°)

117.594

117.952

118.365

118.428

117.962

118.342

117.095

118.321

117.685

117.689

∠S-Sb-S (°)

105.122

104.918

104.681

104.645

104.912

104.694

105.405

104.707

105.070

105.068

r(Bi-S) (Å)

2.525

2.532

2.515

2.519

2.534

2.528

2.557

2.524

2.538

2.548

∠Bi-S-Bi (°)

117.385

117.961

118.369

118.349

118.009

118.275

116.783

118.325

117.497

117.473

∠S-Bi-S (°)

105.241

104.913

104.679

104.690

104.885

104.733

105.581

104.704

105.177

105.191

r(Sb-S) (Å) Sb4S6

MP2

DFT methods

[2.136]

r(As-S) (Å) As4S6

ab-initio methods

2.266

Table 2. Natural atomic charges of M4S6 (Μ = P, As, Sb, Bi) MP2/LANL08(d) CCSD/LANL08(d) M S M S 0.426 -0.284 0.424 -0.283 P4S6 As4S6

0.665

-0.444

0.663

-0.442

Sb4S6

1.073

-0.715

1.077

-0.718

Bi4S6

1.154

-0.769

1.170

-0.780

27

Table 3. Electrostatic potential surface analysis results at CCSD/LANL08(d) level of theory. VvdW

d

Π

2 𝜎𝑡𝑜𝑡

𝜎+2

𝜎−2

(Å3)

(g cm-3)

(kcal mol-1)

(kcal mol-1)2

(kcal mol-1)2

(kcal mol-1)2

𝜈

2 𝜈𝜎𝑡𝑜𝑡

(kcal mol-1)2

P4S6

285.09

1.84

6.12

28.43

24.26

4.17

0.125

3.55

As4S6

306.78

2.67

7.85

38.91

32.38

6.53

0.139

5.44

Sb4S6

350.17

3.22

11.17

59.84

47.81

12.03

0.161

9.61

Bi4S6

360.56

4.74

16.55

111.42

87.05

24.37

0.171

19.04

28

Table 4. NBO chemical bonding and bond ionicity parameter for all M4S6 molecules. (𝝈)

(𝝈)

(𝝅)

|𝐢𝐌𝐒 |

𝝈𝑴𝑺

𝐧𝐌

P4S6

𝜎𝑀𝑆 = 0.635(𝑠𝑝6.93 𝑑 0.14 )𝑃 + 0.772(𝑠𝑝6.00 𝑑 0.07 )𝑆

𝑠𝑝0.59

𝑠𝑝0.39

3𝑝

0.193

As4S6

𝜎𝐴𝑠𝑆 = 0.599(𝑠𝑝9.78 𝑑 0.13 )𝐴𝑠 + 0.800(𝑠𝑝6.53 𝑑 0.06 )𝑆

𝑠𝑝0.38

𝑠𝑝0.38

3𝑝

0.282

Sb4S6

𝜎𝑆𝑏𝑆 = 0.534(𝑠𝑝12.59 𝑑 0.11 )𝑆𝑏 + 0.846(𝑠𝑝6.58 𝑑 0.04 )𝑆

𝑠𝑝0.28

𝑠𝑝0.36

3𝑝

0.430

Bi4S6

𝜎𝐵𝑖𝑆 = 0.518(𝑠𝑝21.00 𝑑 0.07 )𝐵𝑖 + 0.856(𝑠𝑝7.58 𝑑0.04 )𝑆

𝑠𝑝0.15

𝑠𝑝0.30

3𝑝

0.464

M4S6

29

𝐧𝐒

𝐧𝐒

Table 5. Orbital composition of frontier orbitals (%) in M4S6 (M = P, As, Sb, Bi) molecules at CCSD/LANL08(d) Atom P P P P S S S S

Orbital 3pz 3py 3px 3s 3pz 3py 3px 3s

MO-26 6.78 34.32 6.78 10.71 0.04 37.64 0.04 1.46

MO-27 6.78 6.78 34.32 10.71 0.04 0.04 37.64 1.46

MO-28 34.32 6.78 6.78 10.71 37.64 0.04 0.04 1.46

MO-29 10.90 39.82 9.05 0.00 2.26 8.27 1.88 12.72

MO-30 28.94 0.03 30.80 0.00 6.01 0.01 6.39 12.72

As As As As S S S S

4s 4px 4py 4pz 3s 3px 3py 3pz

Sb Sb Sb Sb S S S S

5s 5px 5py 5pz 3s 3px 3py 3pz

11.97 30.80 6.19 6.19 1.08 41.42 0.30 0.30 16.58 6.66 27.69 6.66 0.76 1.41 37.73 1.41

11.97 6.19 6.19 30.80 1.08 0.30 0.30 41.42 16.58 6.66 6.66 27.69 0.76 1.41 1.41 37.73

11.97 6.19 30.80 6.19 1.08 0.30 41.42 0.30 16.58 27.69 6.66 6.66 0.76 37.73 1.41 1.41

0.00 1.01 26.26 37.53 11.61 0.17 4.45 6.37 0.00 0.95 41.45 29.86 9.83 0.10 4.16 3.00

0.00 42.20 16.94 5.65 11.62 7.16 2.88 0.96 0.00 47.22 6.73 18.30 9.83 4.74 0.68 1.84

Bi Bi Bi Bi S S S S

6s 6px 6py 6pz 3s 3px 3py 3pz

12.04 4.00 20.38 4.00 0.52 1.40 55.41 1.40

12.04 4.00 4.00 20.38 0.52 1.40 1.40 55.41

12.04 20.38 4.00 4.00 0.52 55.41 1.40 1.40

0.00 48.00 9.65 14.61 9.10 4.54 0.91 1.38

0.00 0.17 38.54 33.57 9.10 0.02 3.64 3.17

Table 6. Calculated harmonic wavenumbers (cm-1) at MP2(full) level of theory using the LANL08(d) or TZVp basis sets, of the M4S6 cage-like molecules, at Td symmetry. In brackets the calculated intensities are presented [Raman activity in Å4·(amu)-1, IR intensity in km·mol-1]. Intensities and activities less than 0.5 have been set equal to zero. Calculations have been performed at the optimized geometry at the specific level of theory (CCSD(full)/LANL08(d) or MP2(full)/TZVp). Νο

Symmetry

1 2 3 4 5 6 7 8 9 10

T1 (i.a.) E (R) T2 (R,IR) T2 (R,IR) A1 (R) T2 (R,IR) T1 (i.a.) E (R) A1 (R) T2 (R,IR)

P4S6

P4S6 (TZVp)

As4S6

As4S6 (TZVp)

Sb4S6

Bi4S6

139 158 [15] 205 [11,2] 294 [2,4] 350 [76] 390 [1,0] 391 456 [2] 431 [0] 480 [2,81]

143 163 [16] 212 [15,2] 302 [1,5] 362 [69] 404 [1,0] 420 474 [2] 441 [1] 502 [1,84]

106 116 [11] 148 [11,2] 212 [0,6] 244 [29] 314 [0,5] 312 326 [7] 340 [43] 397 [2,75]

107 120 [13] 152 [13,2] 217 [0,7] 250 [21] 331 [0,4] 344 343 [7] 356 [48] 417 [1,66]

79 85 [10] 108 [11,3] 158 [0,6] 181 [16] 271 [1,12] 286 273 [8] 287 [69] 372 [2,110]

70 63 [8] 82 [2,8] 133 [1,9] 133 [9] 245 [1,19] 268 240 [13] 266 [83] 359 [3,102]

assignment β(SMS) + τ(MSMS) + δ(SSSM) β(SMS) + β(MSM) + τ(MSMS) β(SMS) + β(MSM) + δ(SSSM) + τ(MSMS) β(SMS) + β(MSM) + δ(SSSM) + τ(MSMS) ν(M-S) ν(M-S) + τ(MSMS) + δ(SSSM) ν(M-S) + β(SMS) ν(M-S) + β(MSM) β(MSM) + δ(SSSM) ν(MS) + τ(MSMS) + δ(SSSM)

The Greek letters ν, β, δ, τ denote stretching, in plane bending, out of plane bending and torsion modes respectively.