Skip to main content

Calculation of the visible-UV absorption spectra of hydrogen sulfide, bisulfide, polysulfides, and As and Sb sulfides, in aqueous solution

Abstract

Recently we showed that visible-UV spectra in aqueous solution can be accurately calculated for arsenic (III) bisulfides, such as As(SH)3, As(SH)2S- and their oligomers. The calculated lowest energy transitions for these species were diagnostic of their protonation and oligomerization state. We here extend these studies to As and Sb oxidation state III and v sulfides and to polysulfides S n 2-, n = 2–6, the bisulfide anion, SH-, hydrogen sulfide, H2S and the sulfanes, S n H2, n = 2–5. Many of these calculations are more difficult than those performed for the As(iii) bisulfides, since the As and Sb(v) species are more acidic and therefore exist as highly charged anions in neutral and basic solutions. In general, small and/or highly charged anions are more difficult to describe computationally than larger, monovalent anions or neutral molecules. We have used both Hartree-Fock based (CI Singles and Time-Dependent HF) and density functional based (TD B3LYP) techniques for the calculations of absorption energy and intensity and have used both explicit water molecules and a polarizable continuum to describe the effects of hydration. We correctly reproduce the general trends observed experimentally, with absorption energies increasing from polysulfides to As, Sb sulfides to SH- to H2S. As and Sb(v) species, both monomers and dimers, also absorb at characteristically higher energies than do the analogous As and Sb(III)species. There is also a small reduction in absorption energy from monomeric to dimeric species, for both As and Sb III and v. The polysufides, on the other hand, show no simple systematic changes in UV spectra with chain length, n, or with protonation state. Our results indicate that for the As and Sb sulfides, the oxidation state, degree of protonation and degree of oligomerization can all be determined from the visible-UV absorption spectrum. We have also calculated the aqueous phase energetics for the reaction of S8 with SH- to produce the polysulfides, S n H-, n = 2–6. Our results are in excellent agreement with available experimental data, and support the existence of a S6 species.

Introduction

In hydrothermal solutions, As and Sb are often present in appreciable concentration,[1] often occurring in association with Ag, Au and Hg, but the identities of the As and Sb species present are not well understood. In neutral to alkaline sulfidic waters at low temperature, thio- species are believed to predominate. [2] The speciation of Sb in sulfidic solutions has been studied for some time, but new results are still emerging. The main questions concern the oxidation state (III or v), the coordination number and the degree of oligomerization of the species. Typically Sb(III) compounds, which essentially have a 5 s2 lone pair orbital, will be trigonal three-coordinate, while Sb(v) compounds, without the lone pair, will be tetrahedral four-coordinate. By 1990 a consensus seemed to emerge that in alkaline sulfidic solutions Sb existed as Sb(III), based on numerous solubility studies[24] and Raman studies. [5]

However, recent EXAFS studies[6, 7] have presented evidence for the presence of Sb(v) species in such solutions. The Sb-S distances determined by EXAFS were more consistent with those for model compounds with four-coordinate Sb(v) than for those with three-coordinate Sb(III) and the coordination numbers from the model fits to the data were close to 4. Recently Helz and coworkers[8] reported the results of a solubility study for stibnite, Sb2S3, and elemental S in equilibrium with alkaline sulfidic solutions, which could be best interpreted in terms of a number of dimeric species, including the mixed Sb(III,v) and the Sb(v,v) dimers, Sb2S52- and Sb2S62-, which were new species, not previously considered. They also presented visible-UV absorption spectra which showed a broad peak around 4.4 eV, consistent with the limited experimental data available on Sb(v) sulfides. In recent work we have calculated energetics[9] for the formation of such oxidized dimer species which are in good agreement with the experimental data of Helz et al.[8]

We had previously calculated structures, energetics and spectra for various Sb(III) monomers and oligomers,[10] assigning the spectra of Wood[5] to a more protonated Sb(III) dimer than in the original work. At that time procedures recently developed to calculate pKas for such species[11] were not yet available. We also noted in ref. 10 that three-coordinate Sb(III) and four-coordinate Sb(v) had very similar Sb-S stretching frequencies so that information in addition to the Raman spectra was necessary to exclude the presence of Sb(v) species in the solutions studied. Additional Raman spectral data has since been presented for the As-S system,[12] but the spectra seem so complex that assigning species based just on the Raman seems very difficult. Additional information has also recently become available from ion-exchange mass spectro-metry, but only limited information on atom ratios can be obtained using this method. [13]

Recently, UV spectroscopy has been used to study acid dissociation in solution, first for H2S[14] and then for As(OH)3. [15]Although the concentrations of the different species were determined primarily through changes in spectral intensities at energies lower than the absorption maximum (on the low energy side of the band), the maxima themselves were determined for As(OH)3 and AsO(OH)2-, and showed a difference of around 0.5 eV. We calculated the UV absorption spectra for both bare gas-phase As(OH)3, its conjugate base and these same species microsolvated with water. [16] The calculations were also extended to some of the oligomers of As(OH)3 and to related species derived from thioarsenious acid As(SH)3. The calculated energies were in good agreement with experiment and it was clear that both protonation state and degree of oligomerization had observable effects upon the spectrum. This indicated that visible-UV spectroscopy could be a useful new technique for assessing speciation in solutions containing metalloid sulfides, and perhaps for polysulfides as well. It therefore became important to establish whether visible-UV spectra could be accurately calculated for a range of such anionic sulfide species in aqueous solution and whether simple recognizable trends in spectral energy were present, which could be used to determine speciation. Although visible-UV spectra have not been extensively studied for As sulfides, the instruments needed to perform such measurements are readily available. The main impediment to such experimental studies is the paucity of studies calculating and interpreting such data using accurate quantum chemical techniques. This paper represents part of our effort to remedy this deficiency.

A detailed knowledge of As and Sb speciation is important for a number of reasons. First, although several different speciation models may be able to explain a limited set of experimental solubility data, extrapolating into new regimes of concentration, P and T can reveal significant differences in both species concentrations and total element concentrations. Second, thermodynamic models for mineral stability and solubility can only be accurately constructed from experimental data if correct speciations are known. Third, the speciation of an element also influences the interaction of that element with mineral surfaces. For example, to understand the well known association between Au and the As and Sb sulfides[16] it is important to understand both the speciation of As and Sb sulfides and the characteristics of the mineral surfaces. Helz et al.[8] have noted that the III,v and v,v Sb sulfide dimers they have characterized will be anionic, rather than neutral as for As hydroxides, and that they will interact unfavorably with mineral surfaces carrying negative charges, leading to desorption of Sb. This is important since Sb2S3 in the presence of S is soluble enough to exceed the drinking water standards for total Sb concentration.

As discussed in more detail in ref. 10 determination of speciation from solubility data alone is often ambiguous. Even if spectral data such as EXAFS or Raman is also available, determining speciation based on comparison with model compound data is often difficult. Using quantum mechanical methods we can evaluate a number of different properties, including structure, energy and spectral properties for a number of candidate species, and search for the best overall fit to the available experimental data. In this case our primary goal is to determine if the visible-UV spectra of polysulfide and metalloid sulfide species show changes with oxidation state, protonation state or degree and type of oligomerization which are diagnostic of their structure.

Computational methods

We use the methods of Hartree-Fock theory and density functional theory. [18] For all the species considered we have calculated geometries at the second order of Møller-Plesset perturbation theory[19] (MP2) using the polarized SBK basis set. [20] Our previous studies[16] on As(OH)3 and related molecules indicated that the effect of geometry on the visible-UV energies was fairly small. In addition, polarized SBK MP2 is a efficient, medium level correlated method which normally gives very accurate geometries at modest computational cost. We have also directly established the small effect of the method of geometry optimization for the important species S42-···4H2O (vide infra). For the calculation of the visible-UV energies we used both all-electron 6-311 +G(2d,p) bases[21] and the polarized SBK bases and employed both HF and DFT type methods. In a few cases we also have also utilized the large "correlation consistent" basis sets from Dunning's group[22] at the aug-cc-pVDZ and aug-cc-pVTZ (augmented correlation-consistent polarized valence double and triple zeta levels). In evaluating excitation energies we used either the configuration-interaction singles method[23] (CIS), the time-dependent Hartree-Fock method[24] (TD HF), the CIS(D) method[23] or the time-dependent density functional method[2527] (TD DFT) method. The DFT studies have been done mainly with the hybrid B3LYP potential. [28] Analyses of these different methods for calculating excitation energies are given in ref. 25 and 26. Basically, CIS describes the excited state wavefunction at a level comparable to Hartree-Fock, using single excitations from the HF determinant. The TD HF method (also called the random phase approximation, RPA) includes some double excitations, giving a slightly correlated description of both ground and excited states, The CIS(D) method also incorporates some double substitution corrections to the CIS energies, in a size-consistent way. TD DFT includes additional electron correlation through the exchange correlation potential. In ref. 25 it was found for several different molecules that TD DFT using the hybrid B3LYP potential gave the best agreement with experiment, consistently giving excitation energies intermediate between those from TD HF and TD DFT with pure DFT potentials. In our studies[16] on As(OH)3 and related molecules we found the TD HF and TD B3LYP results appeared to bracket the experimental values, with TD B3LYP a few tenths of an eV closer to experiment, but consistently a little too low.

To describe the species in aqueous solution we use both an explicit supermolecule or microhydration approach, in which several water molecules (typically 4–6) are coordinated to the solute at the optimized geometry of the supermolecule, and a polarizable continuum approach. These geometries are similar to those we previously reported[29] for AsO(OH)2-···4H2O, a supermolecule which was used to calculate the vibrational spectrum of the arsenite anion in solution. The geometries of all the supermolecules in the present study were obtained by a similar approach, but utilizing the polarized SBK basis set at the MP2 level. The geometries have been verified to be local energy minima by frequency analysis, but are not necessarily the global minima. These species are in fact just simple approximations to the real hydrated species. We have not systematically studied the effect of varying the number of water molecules in the supermolecule, although we have employed a very large supermolecule with 22 water molecules for S42- (with the geometry optimized at the polarized SBK HF level, not MP2).

Generally better results are obtained for solutes in aqueous environments if the solute is immersed in a polarizable continuum. For the most recent and complete such approach applied to spectral properties see ref. 30. It is also possible to use a "mixed" approach, employing both microhydration and a polarizable continuum. However, in many of the studies done so far with this approach the number of water molecules used has been very small, usually only one or two coordinating to the chromophoric group of the molecule, e.g. the C=O in a study of acetone. [31] We have tried to approach the problem fairly simply yet systematically for the present species, using several waters of hydration (4–6) in the microhydrated species and then immersing this species in a polarizable continuum. We have utilized primarily the COSMO or CPCM version[32] of the polarizable continuum model, although we have also tested the computationally less stable isodensity polarized continuum model. [33] The COSMO solvation approach has been applied both to bare anionic solutes and to the microhydrated species. For the very largest species considered, S42-··· 22H2O, we have employed the ONIOM method,[34] in which different parts of the supermolecule can be treated at different levels of accuracy, using different basis sets or even different quantum methods. We employed the 6-311 +G(2d,p) basis for the S atoms and the 6-31G* basis for the surrounding waters in the CIS ONIOM calculations.

The calculations were performed using the software packages GAMESS,[35] GAUSSIAN94[36] and GAUSSIAN98,[36] on two different clusters of COMPAQ DECStations.

Results and discussion

The main results of our spectral calculations are given in Tables 1 and 2. Table 1 gives results for hydrogen sulfide, bisulfide anion, polysulfide anions and gas-phase sulfanes. Results for As and Sb sulfide monomers and dimers are given in Table 2. Most of the results are obtained using the CIS and CIS COSMO methods with the 6-311+G(2d,p) basis set. Some results using the TD B3LYP method are given in italics. Both free anions and explicitly hydrated anions are considered. We can compare with previous calculations of excitation energies using (and referring to) a number of different methods for H2S and S2H2. [37] Experimental energies where available are given in the last column. The experimental values are taken from ref. 38–42 for H2S, SH-, polysulfides, Sb sulfides and sulfanes, respectively. Calculated oscillator strengths are given in parentheses for some of the compounds, particularly the more symmetric ones, where some transitions are forbidden at the equilibrium geometry.

Table 1 Calculated and experimental values of the lowest energy optical transitions (eV) for hydrogen sulfide, bisulfide anion, poly-sulfide anions and sulfanes, using the 6-311+G(2d,p) basis (TD B3LYP results in italics, CIS oscillator strengths in parentheses)
Table 2 Calculated and experimental lowest energy optical transitions in As and Sb iii and v sulfide monomers and dimers (TD B3LYP results in italics, CIS oscillator strengths in parentheses)

To establish the protonation state expected for the polysulfides and As and Sb sulfides in the solutions of interest, we employ the procedure of ref. 9, in which we calculated gas-phase deprotonation energies at the polarized SBK MP2 level, and hydration energies at the 6-31G* HF level using COSMO. This allows us to calculate an approximate aqueous deprotonation energy, called ΔGaq in ref. 9, which we could then correlate with pKas (of any order for polyprotic acids) using the equation:

pKa= 0.323 ΔGaq - 87.3

The ΔGaq values obtained for S4H-, SbS3(SH)2- and SbS2(SH)2- are 288.1, 287.1 and 298.4 kcal mol-1, respectively, which yield predicted pKas of 5.8, 5.4 and 9.1. This indicates that S4H- and SbS3(SH)2- will be fully deprotonated at neutral pH. Sb(III) species like SbS2(SH)2- will be fully deprotonated above pH 9. These compounds are representative of all the polysulfides, and of all the As and Sb III and v species. Thus, we carried out our absorption energy calculations for the fully deprotonated species.

There are several conclusions which can be drawn based on the data in Table 1:

(1) For H2S and S2H2 our values agree almost as well with experiment as do those from the older, more traditional methods (such as multireference CI), used in ref. 37. The best previous calculated values are 6.37 and 4.98 eV for H2S and S2H2, respectively, while we calculate 6.4 and 5.4 eV, respectively, and the experimental values are 6.4 and 5.0 eV, respectively.

(2) The CIS COSMO results for explicitly hydrated anions are in the best agreement with experiment, although the calculated values tend to be too high by a few tenths of an eV.

(3) The TD B3LYP COSMO results are systematically around 1 eV too low.

(4) Employing COSMO and CIS together always increases the calculated absorption energies compared to CIS alone, but the change is much larger for bare anions than for explicitly hydrated anions.

(5) The effect of either COSMO or explicit hydration is larger for the dianions than for monoanions and neutrals.

(6) The general experimental trend in visible-UV absorption energies of: polysulfides < bisulfide < H2S is correctly reproduced.

(7) The calculations give essentially the same absorption energies for S42- and S4H-, so that distinguishing protonation state in the polysulfides may not be possible.

(8) The nonlinear variation in lowest absorption energy with n for the polysulfide series, S n 2-, seems to be qualitatively reproduced.

(9) For the sulfane series, S n H2, the calculated effect of n on the absorption energy is very small, inconsistent with the values tabulated in ref. 42 (although the spectra are very broad and do not seem to necessarily support the tabulated energy values).

(10) CIS (ONIOM) results for S42- explicitly hydrated by a 22 water molecules are quite similar to the CIS COSMO results for S42-···4H2O, so that larger explicitly solvated clusters could be used to replace continuum solvation (although at considerable additional computational cost).

We also find that the TD COSMO and CIS COSMO results are very similar (although TD requires about twice the computer time of CIS). For example, for S42-···4H2O the lowest excitation energy calculated with TD COSMO is 3.64 eV, only about 0.1 eV lower than with CIS COSMO. Similar close agreement of CIS and TD HF results was seen in our previous study of As(iii) oxo and thio acids. [16] We also find that 6-311+G(2d,p), aug-cc-pVDZ and aug-cc- pVTZ bases also give very similar results (to within 0.1 eV) for S42-···4H2O with the CIS TD method as a test case. Since the aug-cc-pVTZ calculations employ a basis set about twice as large and require about 24 times more computer time than those for aug-cc-pVDZ or 6-311+G(2d,p) bases we therefore have used only the 6-311+G(2d,p) basis for the other species. Therefore, we will from now on quote only the CIS results with the 6-311+G(2d,p) basis for polysulfides.

We have also done a few calculations using the CIS(D) method, for H2S, S32- and S2H2. The changes from the CIS results are small but apparently in the direction to better match experiment. For example the lowest excitation energy in H2S changes from 6.2 to 6.3 eV (exp. 6.4 eV) while that in S2H2 changes from 5.4 to 5.1 eV (exp. 5.0 eV). For S32- the CIS(D) result is 2.7 eV, compared to 3.2 eV with CIS. A more relevant calculation to compare with experimental polysulfide spectra would be CIS(D) with COSMO solvation for S32-···4H2O, but this calculation is presently a bit too demanding of computer time (for S2H2 CIS(D) takes twice as much computer time as CIS).

For S42-···4H2O we have also performed CIS COSMO calculations using the 6-311 +G(2d,p) basis set at four different sets of optimized geometries, obtained using polarized SBK HF, polarized SBK MP2, 6-31G* MP2 and 6-31G* B3LYP methods. The lowest energy transitions calculated for these four different optimized geometries were 3.75, 3.74, 3.70 and 3.65 eV, respectively, indicating that the effect of geometry upon the absorption energy is fairly small and/or that these different methods give very similar optimized geometries.

Based on the data on As and Sb sulfides in Table 2, some additional conclusions can be drawn:

(1) For the As and Sb sulfides, the CIS COSMO results still seem the most reliable, but CIS COSMO and TD B3LYP COSMO results now appear to straddle the very limited experimental data, with CIS COSMO too high and TD B3LYP COSMO too low.

(2) The polarized SBK and the 6-311G(2d,p) bases give very similar results, while the bases with diffuse functions (e.g. 6-311 +G(2d,p)) have a tendency to give energies which are too low, even when explicitly hydrated and stabilized within the polarizable continuum (a similar difficulty with diffuse functions was encountered previously by Tossell for As(OH)3anionic species[16]).

(3) The effects of explicit hydration and COSMO solvation are larger than seen for the polysulfides in Table 1, probably because of the larger charges on the anions.

(4) The neutral acid molecule As(SH)3 shows absorption at higher energy than its anion AsS33- by about 1 eV (in our earlier study As(SH)3 and AsS(SH)2- showed a difference of about half an eV).

(5) For the monomeric species, the calculated energies for oxidation state v are only slightly, but consistently, higher than those for oxidation state III.

(6) The dimeric species have slightly lower absorption energies than the monomers, particularly if we ignore the symmetry forbidden character of the lowest energy transition in Sb2S42- (which restriction would be relaxed away from the equilibrium geometry).

(7) The lowest energy absorption for a mixed bisulfide, polysufide cluster As(S4)(SH) lies between that for S42- and As(SH)3.

As suggested by many researchers, such mixed bisulfide, polysulfide clusters are quite possible and it would probably be worthwhile to investigate their properties more systematically.

We can gain some understanding of the effect that polarizable continuum hydration has on the calculated absorption energies by examining the data in Table 3 for S42-, where we give calculated HOMO and LUMO energies, along with absorption energies calculated with CIS, COSMO and CIS IPCM methods. We see that using COSMO or IPCM increases the HOMO-LUMO gap, Δε, by on the order of 2 eV, since the HOMO is stabilized more than the LUMO by the charge distribution induced in the polarizable continuum. This increase in the HOMO-LUMO gap increases the absorption energy. Note that although the lowest energy transition has a large contribution from the HOMO to LUMO excitation, as discussed in detail in ref. 16 for the As(iii) compounds, there are also other sizable orbital contributions, so that the change in the CIS transition energy is less than the change in the HOMO-LUMO gap.

Table 3 Analysis of changes in UV energies (in eV) between CIS, CIS COSMO and CIS IPCM for S42-(using the 6-311+G(2d) basis)

As shown in Table 4, we can also establish the independence of our CIS COSMO energies from other energies that would be associated with charge-transfer-to-solvent (CTTS) type transitions (such as the ionization energies of the species) and their resemblance to singlet-triplet excitation energies calculated at the HF level, as discussed in a recent study of the CTTS spectra of I- in water. [43] These are essentially test calculations to establish that the CIS COSMO results are a reflection of the electronic structure of the solute and are not describing ionization coupled with acceptance of electrons into an empty orbital of the solvent. That is, the spectra we are considering are not really CTTS spectra. It is sometimes assumed that any solution spectrum which changes as a function of polarity of the solvent is of CTTS nature. This is definitely not the case: transitions which are essentially internal to the solute will be modifed to some extent by changing polarity in the solvent. Of course, the close similarity of CIS COSMO results for the bare and the explicitly hydrated SH- and S42- anions in Table 4 (and for others in the previous tables) also supports this interpretation.

Table 4 Comparison of calculated ionization energies and singlet-triplet excitation energies (evaluated at HF level, including COSMO hydration) and CIS energies (all energies in eV)

It is also clear from the effect of the COSMO solvation on the spectral energies that we can in fact model changes in the absorption spectra which are associated with changes in the dielectric constant of the solvent, which may occur as a result of temperature variation. In Table 5 we show calculated absorption energies for several different species, both free (i.e. dielectric constant of zero) and for continuum solvation with dielectric constants of 38.2 and 78.5 (appropriate to water at 25°C). The value of 38.2 is selected somewhat arbitrarily as a value intermediate between 0 and 75.8 (this is also the value for nitromethane, a common organic solvent) but similar values of water dielectric constants are in fact found for supercritical conditions. It is clear that the absorption energies increases systematically with dielectric constant and that the effect becomes larger for both larger charge magnitudes and smaller sizes of the anions. Thus, this general method could be used to study changes in absorption energy with T in aqueous solution, although it is not yet clear whether the explicit hydration of the species would also need to be modified, along with the dielectric constant.

Table 5 Calculated effect of varying the dielectric constant on the lowest optical transition energies (in eV) for H2S, SH-, S42-···4H2O and SbS33-···4H2O, using the CIS and CIS COSMO methods and the 6-311 + G(2d,p) basis unless otherwise noted (D = dielectric constant)

It would also be desirable to at least put some constraints on the thermodynamic stabilities of the species present in solutions in equilibrium with sulfide, sulfur and metalloid sulfides. We have previously done this for the Sb sulfide species in ref. 9. For this reason we have also directly calculated quantum mechanically the energetics for the formation of various polysulfides, starting from S8 and SH- as reactants, and have compared them with experiment in Table 6. We tabulate gas-phase reaction energies, gas-phase vibrational, translational and rotational contributions to ΔG, the hydration contributions to ΔG, the total ΔG, in solution and the experimental value of ΔG. The calculated free energies are based on polarized SBK MP2 energies and vibrational frequencies for the species involved, with hydration energies evaluated using COSMO. The experimental free energies at 25°C are obtained from the equilibrium constants tabulated by Shea and Helz,[44] based on polysulfide equilibria from Giggenbach. [45] We see that experimental and calculated free energies are in very good agreement, with discrepancies less than 1 kcal mol-1, so long as we describe the reactants as SH- and S8, rather than the rhombic sulfur mentioned in Giggenbach's paper. Since rhombic sulfur and S8 differ in standard enthalpy of formation by almost 25 kcal (mol of S8)-1, changing the reactant to rhombic sulfur would destroy the present agreement of absolute energetics (although trends in free energies in the S n H- series would be unaffected). There is some controversy about the existence of the S6 polyspecies, which is disfavored by Giggenbach but used in one of the models developed by Teder. [46] Our calculations indicate a degree of stability for S6H- similar to that of the earlier members of the polysulfide series.

Table 6 Calculated energetics for the reaction of S8 with SH- to produce polysulfides, with gas-phase energies and vibrational corrections evaluated at the polarized SBK MP2 level, and with COSMO solvation

We have also evaluated the energy for a related reaction of gas-phase molecules, which eliminates the problem of defining the reactants and also eliminates the need to calculate a hydration energy contribution to the reaction energy. This reaction is

3/8 S8(g) + H2S(g) → S4H2(g)

and all the necessary heat of formation data is available in the tabulations of Wagman et al. [47] This is clearly the neutral gas-phase analog of the second equation in Table 6. The experimental value of ΔH is 6.3 kcal mol-1, while the value calculated at the polarized SBK MP2 level for ΔH is 5.8 kcal mol-1, the same excellent level of agreement as for the solution reactions in Table 6. For this gas-phase reaction the calculated ΔG value is 7.6 kcal mol-1 (more positive than the ΔH since the – TΔS term is positive because of the reduction in number of moles of gas). This indicates that the energetics for reaction of S8 to form polysulfide-like species are not greatly changed in going from neutral molecules in the gas-phase to ions in solution, with ΔG values around +8 in the gas-phase for the formation of the n = 4 sulfane and +4 in solution for formation of the n = 4 polysulfide. The good agreement of calculation and experimental for this sulfane reaction indicates that the equilibrium properties of H2S, sulfane systems[48] could also be determined using such methods.

Conclusion

A method has been developed to calculate visible-UV absorption energies for anionic species in aqueous solution which gives both absolute energies and energy trends which are consistent with experiment. The method involves standard CIS calculations with standard large, polarized basis sets on microhydrated species within a polarizable continuum. It is anticipated that the new, more complete polarizable continuum approaches to spectral properties[30] may yield even better results. The basic trends in absorption energy for the different types of species expected in metalloid sulfidic solutions, such as bisulfide, polysulfide and metalloid sulfides, are described correctly. The calculations also predict observable changes in spectral energy for the metalloid sulfides as a function of oxidation state, protonation state and degree of oligomerization, which will provide an additional spectral tool for determining speciation.

References

  1. Ballantyne JM, Moore JN: Geochim Cosmochim Acta. 1988, 52: 475-10.1016/0016-7037(88)90102-0.

    Article  Google Scholar 

  2. Spycher NF, Reed MH: Geochim Cosmochim Acta. 1989, 53: 2185-10.1016/0016-7037(89)90342-6.

    Article  Google Scholar 

  3. Krupp RE: Geochim Cosmochim Acta. 1988, 52: 3005-10.1016/0016-7037(88)90164-0.

    Article  Google Scholar 

  4. Helz GR, Tossell JA, Charnock JM, Pattrick RAD, Vaughan DJ, Garner CD: Geochim Cosmochim Acta. 1995, 59: 4591-10.1016/0016-7037(95)00330-4.

    Article  Google Scholar 

  5. Wood SA: Geochim Cosmochim Acta. 1989, 53: 237-10.1016/0016-7037(89)90376-1.

    Article  Google Scholar 

  6. Mosselmans JFW, Helz GR, Pattrick RAD, Vaughan DJ: Appl Geochem. 2000, 15: 879-10.1016/S0883-2927(99)00080-3.

    Article  Google Scholar 

  7. Sherman DM, Ragnarsdottir KV, Oelkers EH: Chem Geol. 2000, 167: 161-10.1016/S0009-2541(99)00207-7.

    Article  Google Scholar 

  8. Helz GR, Valerio MS, Capps NE: Environ Sci Tech. 2002, 36: 943-10.1021/es011227c.

    Article  Google Scholar 

  9. Tossell JA: Geochim Cosmochim Acta. 2003, in press

    Google Scholar 

  10. Tossell JA: Geochim Cosmochim Acta. 1994, 58: 5093-10.1016/0016-7037(94)90296-8.

    Article  Google Scholar 

  11. Tossell JA, Sahai N: Geochim Cosmochim Acta. 2000, 64: 4097-10.1016/S0016-7037(00)00497-X.

    Article  Google Scholar 

  12. Wood SA, Tait CD, Janecky DR: Geochem Trans. 2002, 3 (4): 31-10.1039/b111453k.

    Article  Google Scholar 

  13. Wilkin RT, Wallschlager D, Ford RG: Geochem Trans. 2003, 4 (1): 1-10.1039/b211188h.

    Article  Google Scholar 

  14. Suleimenov OM, Seward TM: Geochim Cosmochim Acta. 1997, 61 (15): 5187-10.1016/S0016-7037(97)00291-3.

    Article  Google Scholar 

  15. Zakaznova-Iakovleva VP, Seward TM, Suleimenov OM: Spectrophotometric determinatin of the first ionisation constant of H3AsO3 from 25 to 300°C. Steam Water and Hydrothermal Systems. Edited by: P. R. Tremaine, P. G. Gill, D. E. Irish and P. V. Balakrishnan. 2002, NRC Research Press, Ottawa, 694-698.

    Google Scholar 

  16. Tossell JA: Aquat Geochem. 2001, 7: 239-10.1023/A:1015217332249.

    Article  Google Scholar 

  17. Renders PJ, Seward TM: Geochim Cosmochim Acta. 1989, 53: 255-10.1016/0016-7037(89)90378-5.

    Article  Google Scholar 

  18. Jensen F: Introduction to Computational Chemistry. 1999, Wiley, New York

    Google Scholar 

  19. Head-Gordon M, Pople JA, Frisch MJ: Chem Phys Lett. 1988, 153: 503-10.1016/0009-2614(88)85250-3.

    Article  Google Scholar 

  20. Stevens WJ, Krauss M, Basch H, Jansen PG: Can J Chem. 1992, 70: 612-10.1139/v92-085.

    Article  Google Scholar 

  21. Hehre WJ, Radom L, Schleyer PvR, Pople JA: Ab Initio Molecular Orbital Theory. 1986, Wiley, New York

    Google Scholar 

  22. Woon DE, Dunning TH: J Chem Phys. 1993, 98: 1358-10.1063/1.464303.

    Article  Google Scholar 

  23. Foresman JB, Head-Gordon M, Pople JA, Frisch MJ: J Phys Chem. 1992, 96: 135-10.1021/j100180a030. Head-Gordon M, Rico RJ, Oumi M, Lee TJ: Chem. Phys. Lett., 1994, 219: 21

    Article  Google Scholar 

  24. Wiberg KB, Hadad CM, Foresman JB, Chupka WA: J Phys Chem. 1992, 96: 10756-10.1021/j100205a032.

    Article  Google Scholar 

  25. Bauernschmitt R, Ahlrichs R: Chem Phys Lett. 1996, 256: 454-10.1016/0009-2614(96)00440-X.

    Article  Google Scholar 

  26. Stratmann RE, Scuseria GE, Frisch MJ: J Chem Phys. 1998, 109: 8218-10.1063/1.477483.

    Article  Google Scholar 

  27. Van Gisbergen SJA, Groeneveld JA, Snijders JG, Baerends E: J Phys Chem A. 1999, 103: 6835-6844. 10.1021/jp991060y.

    Article  Google Scholar 

  28. Becke AD: J Chem Phys. 1992, 96: 2155-10.1063/1.462066.

    Article  Google Scholar 

  29. Tossell JA: Geochim Cosmochim Acta. 1997, 61: 1613-10.1016/S0016-7037(97)00030-6.

    Article  Google Scholar 

  30. Cossi M, Barone V: J Chem Phys. 2001, 115: 4708-10.1063/1.1394921.

    Article  Google Scholar 

  31. Pappalardo RR, Reguero M, Robb MA, Frisch M: Chem Phys Lett. 1993, 212: 12-10.1016/0009-2614(93)87100-H.

    Article  Google Scholar 

  32. Truong TN, Stefanovich EV: Chem Phys Lett. 1995, 240: 253-10.1016/0009-2614(95)00541-B.

    Article  Google Scholar 

  33. Foresman JB, Keith TA, Wiberg KB, Snoonian J, Frisch MJ: J Phys Chem. 1996, 100: 16098-10.1021/jp960488j.

    Article  Google Scholar 

  34. Svensson M, Humbel S, Froese RDJ, Matsubara T, Sieber S, Morokuma K: J Phys Chem. 1996, 100: 19357-10.1021/jp962071j.

    Article  Google Scholar 

  35. Schmidt MW, et al: J Comput Chem. 1993, 14: 1347-10.1002/jcc.540141112.

    Article  Google Scholar 

  36. Frisch MJ, et al: Gaussian 94 Rev B3 Gaussian Inc. 1995, Pittsburgh, PA, Frisch MJ: et al, Gaussian 98, Rev. A.1, Gaussian Inc., Pittsburgh, PA, 1998

    Google Scholar 

  37. Pericou-Cayere M, Gelize M, Dargelos A: Chem Phys. 1997, 214: 81-10.1016/S0301-0104(96)00300-X.

    Article  Google Scholar 

  38. Herzberg G: Molecular spectra and molecular structure. III. Electronic spectra and electronic structure of polyatomic molecules. 1996, Van Nostrand Reinhold Co., New York, 14-see also ref. 14

    Google Scholar 

  39. Fox M, Hayon E: J Chem Soc Faraday Trans 1. 1979, 75: 1380-10.1039/f19797501380.

    Article  Google Scholar 

  40. Giggenbach WF: Inorg Chem. 1972, 11: 1201-10.1021/ic50112a009.

    Article  Google Scholar 

  41. Kunkely H, Vogler A: Z Naturforsch. 1995, 50b: 1155-

    Google Scholar 

  42. Feher F, Munzner H: Z Anorg Allg Chem. 1963, 96: 1131-

    Google Scholar 

  43. Chen H-Y, Sheu W-S: J Am Chem Soc. 2000, 122: 7534-10.1021/ja000207s.

    Article  Google Scholar 

  44. Shea D, Helz GR: Geochim Cosmochim Acta. 1988, 52: 1815-10.1016/0016-7037(88)90005-1.

    Article  Google Scholar 

  45. Giggenbach WF: Inorg Chem. 1974, 13: 1724-10.1021/ic50137a038.

    Article  Google Scholar 

  46. Teder A: Acta Chem Scand. 1971, 25: 1722-

    Article  Google Scholar 

  47. Wagman DD, Evans WH, Parker VB: The NBS tables of chemical thermodynamic properties. Selected values for inorganic and C1 and C2 organic substances in SI units. 1982, ACS, Washington, DC

    Google Scholar 

  48. Migdisov AA, Suleimenov OM, Alekhin YV: Geochim Cosmochim Acta. 1998, 62: 2627-10.1016/S0016-7037(98)00188-4.

    Article  Google Scholar 

Download references

Acknowledgements

This work was supported by NSF Grant EAR-0001031 and DOE Grant DE-FG02-94ER14467. The COSMO hydration energy calculations and the TD B3LYP calculations were performed using GAUSSIAN98 on the Carnegie Alpha Cluster, which is supported in part by NSF MRI Grant AST-9976645

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to JA Tossell.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Tossell, J. Calculation of the visible-UV absorption spectra of hydrogen sulfide, bisulfide, polysulfides, and As and Sb sulfides, in aqueous solution. Geochem Trans 4, 28 (2003). https://doi.org/10.1186/1467-4866-4-28

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/1467-4866-4-28

Keywords