Your browser doesn't support javascript.
loading
Show: 20 | 50 | 100
Results 1 - 20 de 26
Filter
Add more filters










Publication year range
1.
Chem Soc Rev ; 48(15): 4338, 2019 Jul 29.
Article in English | MEDLINE | ID: mdl-31305818

ABSTRACT

Correction for 'The role of polysulfide dianions and radical anions in the chemical, physical and biological sciences, including sulfur-based batteries' by Ralf Steudel et al., Chem. Soc. Rev., 2019, 48, 3279-3319.

2.
Chem Soc Rev ; 48(12): 3279-3319, 2019 Jun 17.
Article in English | MEDLINE | ID: mdl-31123728

ABSTRACT

The well-known tendency of sulfur to catenate is exemplified by an extensive series of polysulfide dianions [Sn]2- (n = 2-9) and related radical monoanions [Sn]˙-. The dianions can be isolated as crystalline salts with appropriate cations and structurally and spectroscopically characterized. Although the smaller radical monoanions may be stabilized in zeolitic matrices, they are usually formed in solution via disproportionation or partial dissociation of the dianions as well as by electrochemical reduction of elemental sulfur. An understanding of the fundamental chemistry of these homoatomic species is key to unravelling their behaviour in a broad variety of chemical environments. This review will critically evaluate the techniques used to characterize polysulfide dianions and radical anions both in solution and in the solid state, i.e. Raman, UV-visible, EPR, NMR and X-ray absorption spectroscopy, X-ray crystallography, mass spectrometry, chromatography and high-level quantum-chemical calculations. This is followed by a discussion of recent advances in areas in which these anionic sulfur species play a crucial role, viz. alkali-metal-sulfur batteries, organic syntheses, biological chemistry, geochemical processes including metal transport, coordination complexes, atmospheric chemistry and materials science.


Subject(s)
Electric Power Supplies , Sulfides/chemistry , Anions/chemistry , Coordination Complexes/chemistry , Crystallography, X-Ray , Free Radicals/chemistry , Metals, Alkali/chemistry , Quantum Theory
3.
Chemistry ; 19(9): 3162-76, 2013 Feb 25.
Article in English | MEDLINE | ID: mdl-23325664

ABSTRACT

The sodium-sulfur (NAS) battery is a candidate for energy storage and load leveling in power systems, by using the reversible reduction of elemental sulfur by sodium metal to give a liquid mixture of polysulfides (Na(2)S(n)) at approximately 320°C. We investigated a large number of reactions possibly occurring in such sodium polysulfide melts by using density functional calculations at the G3X(MP2)/B3LYP/6-31+G(2df,p) level of theory including polarizable continuum model (PCM) corrections for two polarizable phases, to obtain geometric and, for the first time, thermodynamic data for the liquid sodium-sulfur system. Novel reaction sequences for the electrochemical reduction of elemental sulfur are proposed on the basis of their Gibbs reaction energies. We suggest that the primary reduction product of S(8) is the radical anion S(8)(˙-), which decomposes at the operating temperature of NAS batteries exergonically to the radicals S(2)(˙-) and S(3)(˙-) together with the neutral species S(6) and S(5), respectively. In addition, S(8)(˙-) is predicted to disproportionate exergonically to S(8) and S(8)(2-) followed by the dissociation of the latter into two S(4)(˙-) radical ions. By recombination reactions of these radicals various polysulfide dianions can in principle be formed. However, polysulfide dianions larger than S(4)(2-) are thermally unstable at 320°C and smaller dianions as well as radical monoanions dominate in Na(2)S(n) (n=2-5) melts instead. The reverse reactions are predicted to take place when the NAS battery is charged. We show that ion pairs of the types NaS(2)˙, NaS(n)(-), and Na(2)S(n) can be expected at least for n=2 and 3 in NAS batteries, but are unlikely in aqueous sodium polysulfide except at high concentrations. The structures of such radicals and anions with up to nine sulfur atoms are reported, because they are predicted to play a key role in the electrochemical reduction process. A large number of isomerization, disproportionation, and sulfurization reactions of polysulfide mono- and dianions have been investigated in the gas phase and in a polarizable continuum, and numerous reaction enthalpies as well as Gibbs energies are reported.

4.
J Phys Chem A ; 114(22): 6370-6, 2010 Jun 10.
Article in English | MEDLINE | ID: mdl-20469883

ABSTRACT

Zinc oxide is a semiconductor as well as a catalyst in important industrial processes, and alkylzinc alkoxide species of the type (RZnOR)(n) are precursors for the preparation of nanoscaled zinc oxide particles. In this work, the structures, thermodynamic properties, dipole moments, and molecular vibrations of substituted zinc oxide molecules and cations have been studied by density functional calculations at the B3LYP/6-31+G(2df,p) level. The neutral tetramers (MeZnOMe)(4) and (MeZnO(t)Bu)(4) are cubane-like species of T(d) symmetry in the gas phase, whereas the cation [(MeZnOMe)(4)](+) is approximately of C(3v) symmetry due to a Jahn-Teller distortion. For the neutral trimers (MeZnOMe)(3) and (MeZnO(t)Bu)(3) two structural isomers of almost identical energy were located on the potential energy surfaces: A closo-cluster with eight ZnO bonds is most stable both as isolated molecule and in a polarizable continuum, but a roof-shaped structure with seven ZnO bonds differs in enthalpy by less than 8 kJ mol(-1). In the case of the analogous cation [(MeZnOMe)(3)](+), the isomer with the roof-shaped Zn(3)O(3) skeleton is 31.7 kJ mol(-1) more stable in the gas phase than the cluster isomer. A particularly long Zn-C bond in the most stable cationic tri- and tetramers explains the loss of a methyl group, observed mass spectrometrically after electron impact ionization. The dissociation of the tetramers (MeZnOMe)(4) and (MeZnO(t)Bu)(4) into 4/3 trimers is endothermic and endergonic, but less so for the tert-butyl derivative compared to the all-methyl cluster. This destabilizing effect of tert-butyl substituents on the tri- and tetrameric clusters follows also from the larger internuclear distances between the non-hydrogen atoms in the structures of (MeZnO(t)Bu)(n) (n = 3 and 4) compared to the analogous all-methyl derivatives. Remarkable differences of the atomic charges of the carbon atoms of methyl or tert-butyl groups attached to oxygen have been found.

5.
J Phys Chem A ; 114(12): 4437-45, 2010 Apr 01.
Article in English | MEDLINE | ID: mdl-20218592

ABSTRACT

The radical HSO is an oxidation product of pollutants such as H(2)S and CH(3)SH in Earth's atmosphere. For the first time, the interaction of HSO and its tautomer HOS with single water molecules to yield the hydrates HSO.nH(2)O and HOS.nH(2)O was studied for n = 1-3, applying the high-level G3X(MP2) theory. A large number of structures corresponding to local minima on the potential energy surfaces has been identified. While gaseous HSO is more stable than HOS, the enthalpy diffference between HSO.nH(2)O and HOS.nH(2)O decreases with increasing degree of hydration and becomes practically zero for n = 3. Thus, in aqueous solution as well as in fog and rain droplets, HOS is expected to compete with HSO. The barrier for the tautomerization of HSO to HOS is dramatically lowered by the presence of water molecules since a cyclic transition state allows a concerted proton shift within the system of neighboring hydrogen bonds. The corresponding activation enthalpy of only 73.5 kJ mol(-1) predicted for the transformation of HSO.2H(2)O into HOS.2H(2)O may be compared to the 202 kJ mol(-1) reported for the tautomerization of the unhydrated gaseous HSO/HOS molecules. The impact of water of hydration on the fundamental vibrational modes of HSO and HOS has also been studied. Furthermore, HOS is predicted to dimerize at low temperatures to give two van der Waals molecules with singlet (symmetry C(2)) or triplet configuration (symmetry C(2h)), the latter being more stable than the singlet isomer. The disproportionation of 2HSO to H(2)S and SO(2) is predicted to be exothermic by -263.5 kJ mol(-1). The reaction of HSO with ozone to HSO(2) and O(2) is also strongly exothermic by -274.0 kJ mol(-1) and seems to proceed without any barrier. HOS forms a 1:1 van der Waals complex with O(3); the redox reaction of its two components is calculated as exothermic by -410.9 kJ mol(-1) and results in a rather stable adduct between HOSO and O(2) with the structure of a peroxo isomer of HOSO(3). This unprecedented hydrogen peroxosulfite radical might open a novel route to atmospheric sulfate without the intermediate formation of SO(2) and SO(3).

6.
Phys Chem Chem Phys ; 12(3): 630-44, 2010 Jan 21.
Article in English | MEDLINE | ID: mdl-20066349

ABSTRACT

Certain sulfur bacteria oxidize thiosulfate enzymatically to sulfate, and derivatives of the amino acid cysteine play an important role as intermediates in this process. Since some of the proposed intermediates have so far been of hypothetical nature, we have investigated the structures and thermodynamic properties of more than 60 related derivatives of cysteine (CysH) by high-level quantum chemical calculations both in the gas phase and in a polarizable continuum using the PCM method to simulate an aqueous solution. Most of these molecules and anions were studied for the first time. Especially for the smaller species several conformational isomers of similar energy were identified; their relative stabilities are mainly determined by intramolecular hydrogen bonds. In contrast to the thiolate ion [Cys](-), the gaseous anions [CysS](-), [CysSO(2)](-), [CysSO(3)](-) and [CysSSO(3)](-) are most stable as zwitterions containing an NH(3) rather than an NH(2) group. This result also holds for the polarizable continuum. On the other hand, the related neutral molecules CysH, CysSH and CysSO(2)H are predicted to exist as NH(2) derivatives rather than zwitterions in the gas phase and this connectivity is predicted for CysH and CysSH also in the polarizable continuum. A model molecule of composition C(4)H(7)N(2)O(2)SH (abbreviated as RSH) simulating the structural environment of a cysteine residue within the peptide chain near the corresponding reaction center of the thiosulfate oxidizing enzyme complex "Sox" was used to elucidate the geometry of the proposed reaction intermediates as well as their thermodynamic properties. In the polarizable phase, the S-sulfonate ions [CysSO(3)](-) and [RSSO(3)](-) are predicted to react exothermically with water to the corresponding thiol and hydrogensulfate ions. These results support the proposed mechanism for enzymatic thiosulfate metabolism. Sulfur dioxide and hydrogensulfite anions are predicted to react exothermically and exergonically with thiolate and persulfide anions to give the corresponding S-sulfinate species [RSSO(2)](-) and [RSSSO(2)](-), respectively. The latter ions help to explain the inhibition of certain thiolate based enzymes by aqueous sulfite, disulfite and dithionite anions in sulfur oxidizing microorganisms.


Subject(s)
Bacteria/metabolism , Cysteine/analogs & derivatives , Cysteine/metabolism , Oxidoreductases/metabolism , Quantum Theory , Sulfur/metabolism , Thiosulfates/metabolism , Cysteine/chemistry , Enzyme Inhibitors/chemistry , Enzyme Inhibitors/pharmacology , Models, Chemical , Models, Molecular , Molecular Conformation , Oxidoreductases/antagonists & inhibitors , Peptides/chemistry , Sulfhydryl Compounds/chemistry , Sulfhydryl Compounds/metabolism , Sulfur Dioxide/chemistry , Thermodynamics
7.
J Phys Chem A ; 113(36): 9920-33, 2009 Sep 10.
Article in English | MEDLINE | ID: mdl-19685866

ABSTRACT

The interaction of thiosulfuric acid and its monoanion with up to three water molecules has been studied by density functional and high-level ab initio calculations. More than 40 molecules and anions both as OH and SH tautomers were investigated. The structures in the gas phase as well as in a polarizable continuum were optimized at the B3LYP/6-31G(2df,p) level of theory, whereas G3X(MP2) single-point calculations were applied to obtain enthalpies and Gibbs energies of the gaseous species. In the gas phase, all monoanions of composition [H,S(2),O(3)](-).nH(2)O (n = 0-3) are predicted to be most stable as SH tautomers [HSSO(3)](-). The enthalpies of hydration are -51 +/- 5 kJ mol(-1) per water molecule. In a polarizable phase simulating the dielectric properties of water the OH forms [SSO(2)(OH)](-) of the hydrated and unhydrated monoanions are always most stable. Thiosulfuric acid and its mono- and dihydrates are most stable in the SH/OH form, both in the gas phase and in the polarizable continuum. However, the trihydrate H(2)S(2)O(3).3H(2)O prefers the OH/OH form in the polarizable phase, whereas in the gas phase the SH/OH tautomer represents the lowest minimum structure. The hydration enthalpy of the acid is slightly smaller than predicted for the monoanions. We predict that both thiosulfuric acid and its monoanion exist as equilibrium mixtures of the corresponding tautomers in aqueous solution. The acid decomposition of thiosulfate involves sulfur transfer reactions, but the formerly accepted reaction between two monoanions producing sulfite and [HSSSO(3)](-) ions is endothermic and endergonic both for the naked anions and their trihydrates. Therefore, we propose that the sulfur transfer takes place between [HSSO(3)](-) and H(2)S(2)O(3) producing [HSSSO(3)](-), SO(2), and H(2)O. This reaction is exothermic and exergonic both in the gas phase and in the polarizable phase. From [HSSSO(3)](-) the longer chain sulfane monosulfonate ions [HS(n)SO(3)](-) are formed by a series of sulfur transfer reactions, and these ions eventually split off homocyclic sulfur molecules S(n). The initial decomposition reaction is hindered by SO(2) which reacts with thiosulfate ions to the novel adduct [O(2)SSSO(3)](2-) which seems to be also a key intermediate in the synthesis of trithionate from thiosulfate and SO(2). The implications of these results for the enzymatic thiosulfate metabolism by sulfur bacteria are discussed. The adduct H(2)S.SO(3), an isomer of thiosulfuric acid, has been studied in addition. Its conversion into H(2)S(2)O(3) is predicted to be exothermic, by -39.5 kJ mol(-1).

8.
Chem Asian J ; 3(6): 1026-34, 2008 Jun 02.
Article in English | MEDLINE | ID: mdl-18418826

ABSTRACT

The structures and energetics of eight substituted bis(thiocarbonyl)disulfides (RCS(2))(2), their associated radicals RCS(2)(*), and their coordination compounds with a lithium cation have been studied at the G3X(MP2) level of theory for R = H, Me, F, Cl, OMe, SMe, NMe(2), and PMe(2). The effects of substituents on the dissociation of (RCS(2))(2) to RCS(2)(*) were analyzed using isodesmic stabilization reactions. Electron-donating groups with an unshared pair of electrons have a pronounced stabilization effect on both (RCS(2))(2) and RCS(2)(*). The S-S bond dissociation enthalpy of tetramethylthiuram disulfide (TMTD, R = NMe(2)) is the lowest in the above series (155 kJ mol(-1)), attributed to the particular stability of the formed Me(2)NCS(2)(*) radical. Both (RCS(2))(2) and the fragmented radicals RCS(2)(*) form stable chelate complexes with a Li(+) cation. The S-S homolytic bond cleavage in (RCS(2))(2) is facilitated by the reaction [Li(RCS(2))(2)](+)+Li(+)-->2 [Li(RCS(2))](*+). Three other substituted bis(thiocarbonyl) disulfides with the unconventional substituents R = OSF(5), Gu(1), and Gu(2) have been explored to find suitable alternative rubber vulcanization accelerators. Bis(thiocarbonyl)disulfide with a guanidine-type substituent, (Gu(1)CS(2))(2), is predicted to be an effective accelerator in sulfur vulcanization of rubber. Compared to TMTD, (Gu(1)CS(2))(2) is calculated to have a lower bond dissociation enthalpy and smaller associated barrier for the S-S homolysis.

9.
Chemistry ; 14(3): 919-32, 2008.
Article in English | MEDLINE | ID: mdl-17994599

ABSTRACT

Zinc chemicals are used as activators in the vulcanization of organic polymers with sulfur to produce elastic rubbers. In this work, the reactions of Zn(2+), ZnMe(2), Zn(OMe)(2), Zn(OOCMe)(2), and the heterocubane cluster Zn(4)O(4) with the vulcanization accelerator tetramethylthiuram disulfide (TMTD) and with the related radicals and anions Me(2)NCS(2)(*), Me(2)NCS(3)(*), Me(2)NCS(2)(-), and Me(2)NCS(3)(-) have been studied by quantum chemical methods at the MP2/6-31+G(2df,p)//B3LYP/6-31+G* level of theory. More than 35 zinc complexes have been structurally characterized and the energies of formation from their components calculated for the first time. The binding energy of TMTD as a bidendate ligand increases in the order ZnMe(2)

11.
Chemistry ; 13(2): 502-14, 2007.
Article in English | MEDLINE | ID: mdl-17013961

ABSTRACT

The structures of a large number of isomers of the sulfur oxides S(n)O with n = 4-9 have been calculated at the G3X(MP2) level of theory. In most cases, homocyclic molecules with exocyclic oxygen atoms in an axial position are the global minimum structures. Perfect agreement is obtained with experimentally determined structures of S(7)O and S(8)O. The most stable S(4)O isomer as well as some less stable isomers of S(5)O and S(6)O are characterized by a strong pi*-pi* interaction between S==O and S==S groups, which results in relatively long S--S bonds with internuclear distances of 244-262 pm. Heterocyclic isomers are less stable than the global minimum structures, and this energy difference approximately increases with the ring size: 17 (S(4)O), 40 (S(5)O), 32 (S(6)O), 28 (S(7)O), 45 (S(8)O), and 54 kJ mol(-1) (S(9)O). Owing to a favorable pi*-pi* interaction, preference for an axial (or endo) conformation is calculated for the global energy minima of S(7)O, S(8)O, and S(9)O. Vapor-phase decomposition of S(n)O molecules to SO(2) and S(8) is strongly exothermic, whereas the formation of S(2)O and S(8) is exothermic if n<7, but slightly endothermic for S(7)O, S(8)O, and S(9)O. The calculated vibrational spectra of the most stable isomers of S(6)O, S(7)O, and S(8)O are in excellent agreement with the observed data.

12.
J Org Chem ; 71(25): 9302-11, 2006 Dec 08.
Article in English | MEDLINE | ID: mdl-17137356

ABSTRACT

The homolytic dissociation of the important vulcanization accelerator tetramethylthiuram disulfide (TMTD) has been studied by ab initio calculations according to the G3X(MP2) and G3X(MP2)-RAD theories. Homolytic cleavage of the SS bond requires a low enthalpy of 150.0 kJ mol-1, whereas 268.0 kJ mol-1 is needed for the dissociation of one of the C-S single bonds. To cleave one of the SS bonds of the corresponding trisulfide (TMTT) requires 191.1 kJ mol-1. Me2NCS2* is a particularly stable sulfur radical as reflected in the low S-H bond dissociation enthalpy of the corresponding acid Me2NC(=S)SH (301.7 kJ mol-1). Me2NCS2* (2B2) is a sigma radical characterized by the unpaired spin density shared equally between the two sulfur atoms and by a 4-center (NCS2) delocalized pi system. The ESR g-tensors of the radicals Me2NCSn* (n = 1-3) have been calculated. Both TMTD and the mentioned radicals form stable chelate complexes with a Li+ cation, which here serves as a model for the zinc ions used in accelerated rubber vulcanization. Although the binding energy of the complex [Li(TMTD)]+ is larger than that of the isomeric species [Li(S2CNMe2)2]+ (12), the dissociation enthalpy of TMTD as a ligand is smaller (125.5 kJ mol-1) than that of free TMTD. In other words, the homolytic dissociation of the SS bonds of TMTD is facilitated by the presence of Li+ ions. The sulfurization of TMTD in the presence of Li+ to give the paramagnetic complex [Li(S3CNMe2)2]+ is strongly exothermic. These results suggest that TMTD reacts with naked zinc ions as well as with the surface atoms of solid zinc oxide particles in an analogous manner producing highly reactive complexes, which probably initiate the crosslinking process during vulcanization reactions of natural or synthetic rubber accelerated by TMTD/ZnO.

13.
Chemistry ; 12(33): 8589-602, 2006 Nov 15.
Article in English | MEDLINE | ID: mdl-16953504

ABSTRACT

The vulcanization of rubber by sulfur is a large-scale industrial process that is only poorly understood, especially the role of zinc oxide, which is added as an activator. We used the highly symmetrical cluster Zn(4)O(4) (T(d)) as a model species to study the thermodynamics of the initial interaction of various vulcanization-related molecules with ZnO by DFT methods, mostly at the B3LYP/6-31+G* level. The interaction energy of Lewis bases with Zn(4)O(4) increases in the following order: COCSSH.

14.
J Phys Chem A ; 110(28): 8912-24, 2006 Jul 20.
Article in English | MEDLINE | ID: mdl-16836455

ABSTRACT

Methylzinc alkoxide complexes are precursors for the preparation of nanosized zinc oxide particles, which in turn are catalysts or reagents in important industrial processes such as methanol synthesis and rubber vulcanization. We report for the first time the structures, energies, atomic charges, dipole moments, and vibrational spectra of more than 20 species of the type [(MeZnOR')n] with R' = H, Me, tBu and n = 1-6, calculated by density functional theory methods, mostly at the B3LYP/6-31+G* level of theory. For R' = Me, the global minimum structure of the tetramer (n = 4) is a highly symmetrical heterocubane but a ladder-type isomer is by only 70.9 kJ mol(-1) less stable. The corresponding trimer is most stable as a rooflike structure; a planar six-membered ring of relative energy 13.5 kJ mol(-1) corresponds to a saddle point connecting two equivalent rooflike trimer structures. All dimers form planar four-membered Zn2O2 rings whereas the monomer has a planar CZnOC backbone. A hexameric drumlike structure represents the global minimum on the potential energy hypersurface of [(MeZnOMe)6]. The enthalpies and Gibbs energies of the related dissociation reactions hexamer --> tetramer --> trimer --> dimer --> monomer as well as of a number of isomerization reactions have been calculated. The complexes [(MeZnOMe)n] (n = 1-3) form adducts with Lewis bases such as tetrahydrofuran (thf) and pyridine (py). The binding energy of py to the zinc atoms is about 65% larger than that of thf but is not large enough to break up the larger clusters. The bimolecular disproportionation of [(MeZnOMe)4] with formation of the dicubane [Zn{(MeZn)3(OMe)4}2] and Me2Zn is less endothermic than any isomerization or dissociation reaction of the heterocubane, but for steric reasons this reaction is not possible if R' = tBu. A novel reaction mechanism for the reported interconversion, disproportionation and ligand exchange reactions of zinc alkoxide complexes is proposed.

15.
Phys Chem Chem Phys ; 8(11): 1292-7, 2006 Mar 21.
Article in English | MEDLINE | ID: mdl-16633609

ABSTRACT

The structures and relative stabilities of 15 S3O2 isomers have been investigated by G3X(MP2), CCSD(T)/aug-cc-pVTZ and MRCI/CASSCF calculations. The global energy minimum is a three-membered sulfur ring with two adjacent sulfoxide groups in a trans conformation, i.e. a vic-disulfoxide of C2 symmetry. The SS bond lengths are 2.136 (2x) and 2.354 angstroms at the CCSD(T)/cc-pVTZ level of theory. There is a strong interaction between the pi* orbitals of the two S=O moieties both in the trans and in the almost degenerate cis conformer. The corresponding chain-like singlet and triplet isomers of connectivity OSSSO lie close in energy (ca. 67 kJ mol(-1)) while five-membered and branched four-membered rings are significantly less stable. The structure of S3O2 is in contrast to that of the isoelectronic analogue S5, which exists as a five-membered twisted heterocycle.

16.
Inorg Chem ; 44(24): 8908-15, 2005 Nov 28.
Article in English | MEDLINE | ID: mdl-16296846

ABSTRACT

The potential energy hypersurfaces (PESs) of heptasulfur (S7) and of [LiS7]+ have been investigated by ab initio molecular orbital calculations at the G3X(MP2) level of theory. Besides the chair-like seven-membered ring (1a) as the global minimum structure, eight local minimum structures and one transition state have been located on the PES of S7. The barrier for pseudorotation of 1a is only 5.6 kJ mol(-1). The boat-like S7 ring (1b) is 12.1 kJ mol(-1) less stable than 1a, followed by three isomers of connectivity S6=S and four open-chain isomers. On the basis of multireference calculations at the MRCI(4,4)+Q/6-311G(d) level, the most stable open-chain form of S7 is a triplet of relative energy 133.1 kJ mol(-1). Thus, the reaction energy (deltaE0) for the ring opening of 1a is 133.1 kJ mol(-1), halfway between those of the highly symmetrical rings S6 and S8. Because of their strong multireference characters, the stabilities of the biradicalic singlet chains are significantly overestimated by the single-reference-based G3X(MP2) method. The calculated vibrational spectrum of 1a is in good agreement with experimental data. The various isomers of S7 form stable complexes with Li+ with coordination numbers of 1-4 for the metal atom and binding energies in the range of -93.8 to -165.7 kJ mol(-1). A total of 15 isomeric complexes have been located, with 13 of them containing cyclic ligands. The global minimum structure (2a) is composed of 1a, with the Li+ cation linked to the four negatively charged sulfur atoms (symmetry C(s)). The corresponding complex 2c containing the ligand 1b is by 23.4 kJ mol(-1) less stable than 2a, and a bicyclic crown-shaped LiS7 cation (2e) is by 34.9 kJ mol(-1) less stable than 2a. Even less stable are four complexes with the branched S6=S ligand. SS bond activation by polarization of the valence electrons takes place on coordination of Li+ to cyclo-S7 (1a).

17.
Chem Commun (Camb) ; (29): 3712-4, 2005 Aug 07.
Article in English | MEDLINE | ID: mdl-16027920

ABSTRACT

Higher-level ab initio calculations showed that the global energy minimum for both S3O and S3O.+ is a branched, three-membered ring, not an open chain form.

18.
Inorg Chem ; 44(9): 3007-9, 2005 May 02.
Article in English | MEDLINE | ID: mdl-15847403

ABSTRACT

According to ab initio MO calculations at the G2(MP2) level of theory, branched isomers of dialkyl triselenides, R-Se(=Se)-Se-R (1; R = Me, Et), are less stable by more than 60 kJ mol(-1) than the isomeric unbranched chains R-Se-Se-Se-R (2). Therefore, species 1 cannot be generated in substantial concentrations under equilibrium conditions at moderate temperatures, as has recently been claimed by Meja and Caruso (Inorg. Chem. 2004, 43, 7486). Alternatively, the isomeric CH3-Se-CH2-Se-Se-Et (3) can be considered to explain the reported gas chromatograms and mass spectra previously assigned to Et-Se(=Se)-Se-Et (1b). However, the isomerization 2b --> 3 is also endothermic, by deltaG(o)298 = 63 kJ mol(-1). The isomeric selenols HSe-C2H4-Se-Se-Et (4) and CH3-CH(SeH)-Se-Se-Et (5) are also less stable than 2b (by ca. 56 kJ mol(-1)), but 4 is another candidate to explain the mass spectrum formerly assigned to 1b. The calculated structures of 1-5 are reported.

19.
Chemistry ; 11(4): 1281-93, 2005 Feb 04.
Article in English | MEDLINE | ID: mdl-15627950

ABSTRACT

Complex formation between gaseous Li+ ions and sulfur-containing neutral ligands, such as H2S, Me2Sn (n = 1-5; Me = CH3) and various isomers of hexasulfur (S6), has been studied by ab initio MO calculations at the G3X(MP2) level of theory. Generally, the formation of LiS(n) heterocycles and clusters is preferred in these reactions. The binding energies of the cation in the 29 complexes investigated range from -88 kJ mol(-1) for [H2SLi]+ to -189 kJ mol(-1) for the most stable isomer of [Me2S5Li]+ which contains three-coordinate Li+. Of the various S6 ligands (chair, boat, prism, branched ring, and triplet chain structures), two isomeric complexes containing the S5==S ligand have the highest binding energies (-163+/-1 kJ mol(-1)). However, the global minimum structure of [LiS6]+ is of C(3v) symmetry with the six-membered S(6) homocycle in the well-known chair conformation and three Li--S bonds with a length of 256 pm (binding energy: -134 kJ mol(-1)). Relatively unstable isomers of S6 are stabilized by complex formation with Li+. The interaction between the cation and the S6 ligands is mainly attributed to ion-dipole attraction with a little charge transfer, except in cations containing the six sulfur atoms in the form of separated neutral S2, S3, or S4 units, as in [Li(S3)2]+ and [Li(S2)(S4)]+. In the two most stable isomers of the [LiS6]+ complexes, the number of S--S bonds is at maximum and the coordination number of Li+ is either 3 or 4. A topological analysis of all investigated complexes revealed that the Li--S bonds of lengths below 280 pm are characterized by a maximum electron-density path and closed-shell interaction.


Subject(s)
Lithium/chemistry , Models, Chemical , Organometallic Compounds/chemistry , Sulfur Compounds/chemistry , Cations/chemistry , Computer Simulation , Ligands , Models, Molecular , Quantum Theory , Thermodynamics
20.
J Chem Phys ; 121(12): 5899-907, 2004 Sep 22.
Article in English | MEDLINE | ID: mdl-15367018

ABSTRACT

High-level ab initio molecular orbital calculations were employed to explore the potential energy hypersurface of hexasulfur, S(6). Twelve isomeric structures of S(6) have been identified: two unbranched rings (chair and boat), one trigonal prism of D(3h) symmetry, two singly branched rings (S(5)double bondS), three triplet chains, one singlet chain, and three doubly branched rings (Sdouble bondS(4)double bondS). The prism structure is essentially a cluster of three S(2) molecules connected via a six-center pi(*)-pi(*)-pi(*) interaction. It is by 51 kJ mol(-1) less stable than the lowest-energy chair form. The reactions to generate the boat, the prism, and the singly branched isomers from the chair form are predicted to have lower barriers than the ring opening reaction of cyclo-S(6), which requires an activation energy of 149 kJ mol(-1). The prism and singly branched isomers are found to be more reactive species than the chair form and they are potential sources of S(2) in chemical reactions involving elemental sulfur.

SELECTION OF CITATIONS
SEARCH DETAIL
...