Abstract
The efficient protocol for estimation of gas-phase enthalpies of formation developed previously for C, H, O, N, and F elements was extended to sulfur. The protocol is based on local coupled-cluster with single, double, and perturbative triple excitations (CCSD(T)) approximation and allows rapid evaluation of compounds with sizes computationally prohibitive to canonical CCSD(T) using quadruple-zeta basis sets. As a part of model development, a comprehensive review and critical evaluation of experimental data was performed for 87 sulfur-containing organic and inorganic compounds. A compact model with only three empirical parameters for sulfur introduced to address the effects beyond frozen-core CCSD(T) was developed. The model exhibits about 2 kJ·mol−1 standard deviation over a set of experimental values for a diverse collection of sulfur-containing compounds. The complete basis set version of the model demonstrates a similar performance and requires only one empirical parameter. Multiple problems with the existing experimental data were identified and discussed. In addition, a lack of reliable data for certain important classes of sulfur compounds was found to impede the model generalization and confident performance assessment.
Graphical Abstract

Introduction
Enthalpy of formation represents an important thermochemical property with numerous practical applications.1 As such, its efficient and accurate estimation is of great interest for process modeling and design,2 as well as for validation of existing and newly produced experimental data.3 Ab initio methods have become an essential resource for prediction of enthalpies of formation;4 yet, until recently, the theory levels required to achieve the accuracy needed by applications were computationally prohibitive for large-scale and routine processing of compounds of practical interest, especially in terms of size. However, with recent availability of local coupled-cluster methods5–13 accelerated with the “resolution of identity” (density fitting) procedures,14–16 accurate approximations of coupled-cluster with single, double, and perturbative triple excitations (CCSD(T)) approach became practical even for moderate computational resources. With these methods, a medium-sized compound can be processed in a matter of a few hours.8,17,18
In our previous work, we developed a protocol based on local CCSD(T) that allows ideal-gas ΔfH° estimation for closed-shell compounds composed of C, H, O, or N atoms with the accuracy competitive with that of a typical experiment.17,18 Later, this protocol was extended to include fluorine.19 Within this approach, ideal-gas enthalpy of formation is estimated as
| (1) |
where E is the total electronic energy, ZPVE is the zero-point vibrational energy, and is the thermal correction from (0 to 298.15) K. The summation in the last term of Eq. 1 is performed over all atomic types present in the compound; ni is the ith type count, and hi is the type-specific constant. Mathematically, Eq. 1 is equivalent to determination of ΔfH° from atomization energy and hi can be formally defined using computed atomic electronic energies, reference enthalpies of formation, and reference enthalpy changes for individual atomic species. Direct use of atomic species as references is problematic, both in terms of implementation within the local coupled-cluster framework and the overall level of theory required for accurate description. For example, Klippenstein et al. recently suggested the use of molecular reference species instead.20 Here, we determine hi empirically, similar to earlier atom-equivalent proposals,21,22 by fitting to thoroughly vetted experimental data. Once the effective atomic type enthalpies are established, the protocol does not require any auxiliary experimental data as in the approaches based on enthalpies of reactions,23 while approximating the enthalpies of reactions of compound in question and the compounds from the data set used to determine hi. Additional benefit of this semi-empirical approach is the flexibility to account for the effects beyond frozen-core local CCSD(T) by introducing different atomic types for the same element (based, e.g., on its bonding patterns) given the theoretical or statistical justification. Previously, we used two different types for carbon18 as was necessitated by the statistical analysis of experimental data.
The success of the approach based on Eq. 1, to a large extent, rests on the quality of the experimental data used to obtain hi, and their critical evaluation represents an essential and vital part of the model development. For the CHON-compounds,17,18 each data point in the diverse data set used for the parameterization of the method had multiple and consistent experimental verifications. This was not the case for fluorinated organics when only limited amount of data was available with multiple experimental problems identified.19 As a result, the uncertainties of predictions were higher as compared to those for the CHON systems, even after exhaustive critical evaluation of the original experimental sources.
In this work, we extend the protocol to a third-row element, sulfur. This poses a number of challenges. First, the issue of the experimental data needs to be addressed. The experimental enthalpies of formation for CHON molecules are normally determined from combustion experiments in a static bomb,24 while the measurements for S-containing compounds require a rotating-bomb calorimeter. This technique have been used only by a few groups. Thus, the experimental ΔfH° values are only available for less than 400 sulfur-containing compounds as compared to several thousands for CHON compounds. These experiments are extremely time- and resource-consuming, and, consequently, are almost never independently verified. In other words, “a single measurement from a single research group” represents the vast majority of the existing data. Second, there are also theoretical challenges. In contrast to the previously considered elements, sulfur has low-lying d-orbitals available to hybridization and exhibits multiple accessible oxidation states.25 Limited high-level theory analysis for compounds with S in high oxidation states (i.e., SO2 and SO3) suggests that the contributions beyond frozen-core CCSD(T) (e.g., scalar relativistic and spin-orbital corrections) differ substantially from those with S(II).26,27 For example, the reported scalar relativistic corrections to atomization energy for H2O, H2S, SO2, and SO3 are −1.1, −1.6, −3.3, and −7.4 kJ·mol−1, respectively.27
It is, therefore, expected that the protocol described by Eq. 1 would require multiple atomic types or other corrections to describe S-containing compounds with target accuracy. Compounds with S in high oxidation states were also shown to present difficulties for popular budget composite methods.28 In the following, we address the challenges outlined above. We conduct a comprehensive critical analysis of the available experimental data and generate the data set for parametrization of Eq. 1 for S-containing closed-shell compounds. As a result, we produce a compact semi-empirical parameterization for sulfur that is consistent with our previous work on CHONF systems,17–19 i.e., without resorting to explicit introduction of ab initio-based corrections beyond frozen-core CCSD(T).
Computational Details
The computations followed the aLL5 protocol from Ref. 18 that was found to be a reasonable compromise between the accuracy and the computational cost on a typical multi-core computer system. A brief description of the protocol follows; for further details, the reader is referred to the original publication. 18 The initial optimization and computation of vibrational frequencies were performed at B3LYP-D3(BJ)/def2-TZVP level. Separate vibrational frequency scaling factors were used for calculation of zero-point vibrational energies (ZPVE) and thermal correction to the enthalpy, . For ZPVE, a single scaling factor of 0.990 was used. For rigid rotor/harmonic oscillator-approximated , factors of 0.960 and 0.985 were applied to hydrogen stretches and all other modes, respectively. The geometries used for local coupled-cluster calculations were optimized with the density-fitted (resolution-of-identity) second-order Møller-Plesset perturbation theory (DF-MP2). The single-point energy calculations were performed with the 2016 version of LCCSD(T) of Kállay et al.11,12 In both calculations, aug-cc-pVQZ basis set was used for all atoms, except sulfur. For sulfur, we utilized aug-cc-pV(Q+d)Z,29 a revised version augmented with tight d-functions to address core polarization effects for third-row atoms.30,31 DFT calculations were performed with Gaussian 16,32 DF-MP2 was done with Psi4 v1.3.2,33,34 and LCCSD(T) was carried out with MRCC (release of February 9, 2019).35,36 All correlated calculations used frozen-core approximation.
For LCCSD(T) extrapolation to complete basis set (CBS), additional LCCSD(T) computations were carried out with quintuple-zeta basis sets, aug-cc-pV(5+d)Z for sulfur and aug-cc-pV5Z for other atoms. For each molecule, these computations were performed only for the conformer with the lowest LCCSD(T) energy obtained with the quadruple-zeta basis set. The SCF energy contribution was extrapolated using the Karton-Martin modification37 of Jensen’s extrapolation formula:38
| (2) |
where
| (3) |
and L = 5.
The CCSD(T) correlation energy contribution was extrapolated with the equation
| (4) |
which, to a large extent, is based on empirical observations (see, e.g., Ref. 39).
As discussed below, additional corrections were evaluated for two selected cases, dimethyl sulfoxide and dimethyl sulfone. The core-valence contributions to atomization energies were computed at the canonical CCSD(T)/aug-cc-pwQZ theory level. The scalar relativistic contributions were found from the differences between the non-relativistic CCSD(T)/aug-cc-pVQZ and the second-order Douglas-Kroll-Hess40,41 CCSD(T)/aug-cc-pVQZ-DK calculations. These calculations were also performed with Psi4. The computations for atoms were conducted using UHF determinants, while atomic spin-orbit coupling energies were calculated using atomic energy levels.
Previously, our CHON training data set was chosen to minimize the conformational contributions to the resulting ΔfH°.17 Here, due to the insufficient amount of experimental data for rigid molecules, those with multiple conformations had to be considered. As previously,18,19 to account for multiple conformations, we adopted the model that assumes the ideal-gas equilibrium mixture of individual conformers with the entropy component of the standard Gibbs energy computed using the same rigid rotor/harmonic oscillator approximation as was used for terms. Enthalpy of formation for a given compound was computed as the Gibbs-energy average for the conformer population.
The following CHNOF parameters in Eq. 1 established previously18,19 were fixed: h(C saturated or aromatic) = −99910.32 kJ·mol−1, h(C unsaturated) = −99909.44 kJ·mol−1, h(H) = −1524.23 kJ·mol−1, h(O) = −197138.05 kJ·mol−1, h(N) = −143612.32 kJ·mol−1, and h(F) = −26711.75 kJ·mol−1.
As discussed above, the effective atomic enthalpy of sulfur will depend on the oxidation state. Additional problems may arise from sulfur-specific bonding configurations. For example, the S–S bond is a σ bond with a significantly decreased π antibonding compared to O-O due to the increased bond length; its description poses significant challenges even for modern high-level methods.42 Direct evaluation of potentially needed high-level corrections would make the protocol more complex and may affect its efficiency. Instead, we propose a set of empirical corrections. Empirical group- and bond-additivity corrections were successfully used in the past, with applications ranging from semi-empirical methods43 to MP444 and G3-theory composite methods.45 The initial empirical model for the effective enthalpy of a sulfur atom was set to accommodate all anticipated patterns, with the intent to simplify its form based on statistical data analysis. Specifically, the effective enthalpy of a sulfur atom is defined as a base value, h(S), with the following corrections applied: S(IV), Δh(S(IV)); S(VI), Δh(S(VI)); S participating in a double bond, ; aromatic S, Δh(S(arom)); S participating in the S–S bond, ; S forming an S–H bond, Δh(SH). For example, the effective enthalpy of a sulfur atom in H2SO4 is defined as . In H2S2, , for each S atom. This introduces seven potential empirical parameters to be established or eliminated.
Experimental data
To obtain a few kJ·mol−1 uncertainty in the enthalpy of formation, the uncertainty of the experimental combustion energy should be as low as a few hundredth percent. This requires state-of-the-art instruments, significant human expertise, and a multistep procedure for reduction of the experimental data to the standard conditions and T = 298.15 K. The current data reduction methodology was established in the mid-1950s,24 and the data published prior to 1970 were revised by Cox and Pilcher.46 In the subsequent analysis, we will use these revised values for the pre-1970 reports on sulfur compounds.
The values of auxiliary quantities used for data reduction have been significantly improved over the last six decades. However, in most studies, the 1956 recommendations24 are still used. Recalculation of all experimental values with the best available recommendations47 goes beyond the scope of this work, and it is not clear a priori whether the changes will be notable. Therefore, for all results published after 1970, we used the enthalpies of formation from the original analysis unless errors in data reduction were found.
Several dozens of sulfur compounds with sufficiently low repeatabilty-based expanded uncertainty (0.95 level of confidence) of the enthalpies of formation have been identified (Tables 4–9). Their chemical structures are shown in Figure 1. Only the compounds with a flexible backbone containing up to 5-6 heavy atoms are considered to avoid dealing with significant conformational ambiguity.
Table 4:
Thermochemical properties of thiols at T = 298 K
| Chemical Name | CASRN | Phase at T | Referencesa | |||
|---|---|---|---|---|---|---|
| Training set | ||||||
| methanethiol (1.1) | 74-93-1 | lc | −(22.3 ± 0.7) | 98; 48 | −0.5 | −0.9 |
| ethanethiol (1.2) | 75-08-1 | l | −(46.0 ± 0.6) | 99; 48 | −0.9 | −1.2 |
| 1-propanethiol (1.3) | 107-03-9 | l | −(67.7 ± 0.7) | 100; 48 | −0.8 | −0.9 |
| 2-propanethiol (1.4) | 75-33-2 | l | −(75.9 ± 0.7) | 100; 101 | −1.3 | −1.0 |
| 2-butanethiol (1.5) | 513-53-1 | l | −(96.6 ± 0.8) | 102; 103 | −0.7 | −0.6 |
| 2-methyl-2-propanethiol (1.6) | 75-66-1 | l | −(109.3 ± 0.9) | 102; 104 | −0.9 | −0.4 |
| cyclopentanethiol (1.7) | 1679-07-8 | l | −(47.7 ± 0.8) | 105 | −1.6 | −1.4 |
| benzenethiol (1.8) | 108-98-5 | l | 112.7 ± 0.8 | 106; 106: 107 | 0.5 | −0.6 |
| Testing set | ||||||
| 1-butanethiol (1.9) | 109-79-5 | l | −(87.8 ± 1.2) | 102; 108,109 | −1.3 | −1.2 |
| 2-methyl-1-propanethiol (1.10) | 513-44-0 | l | −(96.9 ± 0.8) | 102; 110 | 0.1 | 0.4 |
| cyclohexanethiol (1.11) | 1569-69-3 | l | −(95.7 ± 0.9) | 56; 111,112 | −2.2 | −1.6 |
| phenylmethanethiol (1.12) | 100-53-8 | l | 93.5 ± 1.2 | 56: 113; 114 | −2.9 | −2.7 |
| 1,2-ethanedithiol (1.13) | 540-63-6 | l | −(9.0 ± 1.2) | 115 | −2.8 | −3.2 |
Sources of combustion or reaction energies used to derive the condensed-phase and sources of enthalpies of sublimation or vaporization are separated by semicolon. Unused sources (if any) are listed after colon for each field;
;
at bomb conditions
Table 9:
Thermochemical properties of problematic compounds at T = 298 K
| Chemical Name | CASRN | Phase at T | Referencesa | |||
|---|---|---|---|---|---|---|
| Reported after 2013 and included in the testing set (see Tables 6 and 7) | ||||||
| 4,4’-disulfanediyldianiline (3.16) | 722-27-0 | cr | 240.5 ± 5.1 | 58 | 7.4 | 8.0 |
| 2-methylbenzothiazole (3.19) | 120-75-2 | l | 140.5 ± 2.8 | 76 | 14.3 | 14.1 |
| 5-fluoro-2-methylbenzothiazole (3.20) | 399-75-7 | l | −(47.0 ± 4.3) | 94 | 9.1 | |
| 1,3-dihydro-5-methoxy-2H-benzimidazole-2-thione (4.8) | 37052-78-1 | cr | 19.6 ± 2.8 | 62 | 5.2 | 5.9 |
| 5-amino-1,3-dihydro-2H-benzimidazole-2-thione (4.10) | 2818-66-8 | cr | 174.6 ± 3.7 | 62 | 6.6 | 6.1 |
| Not used in model development | ||||||
| phenoxathiin (6.1) | 262-20-4 | cr | 126.4 ± 1.9 | 77; 77,168,169 | −6.0 | −7.3 |
| 1,3,5-trithiane (6.2) | 291-21-4 | cr | 84.6 ± 2.6 | 53; 53,54 | −8.2 | −12.0 |
| 1,3-dihydroimidazole-2-thione (6.3) | 872-35-5 | cr | 164.9 ± 1.7 | 63 | −31.9 | −32.2 |
| dimethyl sulfone (6.4) | 67-71-0 | cr | −(370.6 ± 1.7) | 64,78; 79: 81 | −8.4 | −7.0 |
| dimethyl sulfite (6.5) | 616-42-2 | l | −(483.3 ± 2.1) | 82 | 5.5 | 5.2 |
| dimethyl sulfate (6.6) | 77-78-1 | l | −(686.6 ± 2.1) | 82 | −8.4 | −8.6 |
| benzenesulfonamide (6.7) | 98-10-2 | cr | −(208.9 ± 2.3)c | 83 | −5.3 | −3.6 |
| cr | −(231.8 ± 3.0) | 84; 83 | 17.6 | 19.3 | ||
| thiacyclohexane 1,1-dioxide (6.8) | 4988-33-4 | cr | −(394.8 ± 1.5) | 85 | −9.8 | −6.7 |
Sources of combustion or reaction energies used to derive the condensed-phase and sources of enthalpies of sublimation or vaporization are separated by semicolon. Unused sources (if any) are listed after colon for each field;
, kJ·mol−1;
revised using the NIST Combustion Calorimetry Tool47
Figure 1:

Chemical structures and identifiers of compounds considered in this work. The compound identifiers correspond to those in Tables 4–9.
The list of laboratories having the largest contributions to the field includes NIPPR (Bartesville, Oklahoma), University of Lund (Sweden), Queen’s University of Belfast (United Kingdom), University of Porto (Portugal), Institute of Physical Chemistry “Rocasolano” (Madrid, Spain), and Benemérita Universidad Autónoma de Puebla (Puebla, Mexico).
Enthalpies of vaporization or sublimation, which are required to convert the experimental condensed-phase enthalpies of formation to the gas-phase values, are often available from several works. For these properties, either weight-averaged or critically evaluated48 experimental values were used.
After preliminary analysis of the experimental data, some values were excluded from the consideration because they ended up having large uncertainties. The remaining data were split into two parts, a training set and a testing set. The former had about 60 % of the considered organic and all inorganic compounds. The testing set primarily contained the results published after 2013 even if their repeatability-based uncertainty exceeded the threshold of 4 kJ·mol−1.
Results and Discussion
In this section, small datasets are used to evaluate the effect of various factors on the computed results. This includes contributions beyond frozen-core CCSD(T), local approximation, and CBS extrapolation. This information gives one a better understanding of the accuracy potentially accessible with the protocol used. Then, a brief discussion of the available data for different groups of sulfur compounds is given. Finally, these results and conclusions are used to derive the effective enthalpies of sulfur, analyze performance of the protocol, and identify the problems and challenges with the experimental data.
Previously,18 we analyzed energy contributions to atomization energies beyond frozen-core CCSD(T), ΔEhoc, and contributions due to local CCSD(T) approximations, ΔEloc, for CHON-containing molecules. These contributions were found to be well described by the atom-equivalent additivity approximation,
| (5) |
where, as in Eq. 1, ni is atomic type count and ei is an empirical constant obtained from regression against ΔEX data. The superscript “X” in the above notations is a placeholder used to distinguish the sources of contributions. In the following sections, we evaluate statistics of for the CHOS-containing compounds using a combination of data available from the literature and generated in this study. In addition, we analyze the effect of CBS extrapolation on the predictions.
Energy contributions beyond frozen-core CCSD(T)
The considered dataset is presented in Table 1. Most selected compounds have the W4.4, W4.2, or W4 results49 available. For H2SO4, the core-valence, scalar relativistic, and spin-orbit contributions to the atomization energies are available50 and included in the analysis. We also computed the latter contributions for two organic compounds, dimethyl sulfoxide and dimethyl sulfone. It was necessary to verify that the behavior of organic S(IV) and S(VI) compounds is similar to that of the inorganic counterparts and to clarify inconsistencies in the experimental data discussed below.
Table 1:
Sum of contributions beyond frozen-core CCSD(T) ΔEhoc), post-CCSD(T) contributions (ΔEpost), contributions other than post-CCSD(T) to atomization energy (ΔEhoc – ΔEpost), and their deviations from the atom-equivalent additivity, kJ·mol−1
| compound | protocol | ΔEhoc | dev.a | ΔEpost | dev.a | ΔEhoc – ΔEpost | dev.a |
|---|---|---|---|---|---|---|---|
| hydrogen | W4.4 | 0.08 | −0.04 | 0.00 | −0.11 | 0.08 | 0.06 |
| water | W4.4 | −0.12 | −0.71 | −0.08 | 0.19 | −0.04 | −0.90 |
| methane | W4.4 | 4.48 | −0.13 | 0.04 | −0.17 | 4.44 | 0.04 |
| ethyne | W4.4 | 9.50 | −0.93 | 0.42 | −0.35 | 9.08 | −0.57 |
| carbon monoxide | W4.4 | 2.64 | 0.75 | 0.38 | −0.08 | 2.26 | 0.83 |
| hydrogen sulfide | W4.4 | −2.18 | 0.97 | 0.29 | 0.76 | −2.47 | 0.21 |
| ethene | W4.2 | 8.58 | 0.03 | 0.08 | −0.13 | 8.49 | 0.16 |
| methanal | W4.2 | 3.31 | 0.12 | 0.33 | −0.14 | 2.97 | 0.27 |
| carbon dioxide | W4.2 | 3.77 | −1.26 | 0.25 | 0.26 | 3.51 | −1.51 |
| hydrogen peroxide | W4.2 | −0.38 | −1.33 | 1.00 | −0.69 | −1.38 | −0.65 |
| ethane | W4 | 7.95 | 0.70 | −0.46 | 0.31 | 8.41 | 0.39 |
| sulfur dioxide | W4 | −1.80 | −1.20 | 1.76 | −0.17 | −3.56 | −1.03 |
| sulfur trioxide | W4 | −6.65 | 2.77 | 1.17 | 0.63 | −7.82 | 2.15 |
| carbon oxide sulfide | W4 | 1.30 | 0.84 | 1.17 | 0.29 | 0.13 | 0.55 |
| carbon disulfide | W4 | 1.72 | 0.05 | 2.22 | 0.19 | −0.50 | −0.14 |
| disulfur monoxide | W4 | −1.63 | −1.74 | 3.47 | −0.94 | −5.10 | −0.80 |
| sulfuric acid | b | −6.78 | 0.16 | ||||
| dimethyl sulfoxide | this work | 5.01 | 0.29 | ||||
| dimethyl sulfone | this work | 2.52 | 1.70 | ||||
| Standard deviation | 1.28 | 0.49 | 0.98 |
Additive value minus the computed value;
Ref. 50
As follows from the results in Table 1, deviations from additivity for the post-CCSD(T) contributions do not exceed ±1.0 kJ·mol−1, which is acceptable for the expected uncertainty of ≈3 kJ·mol−1. The deviations are significantly higher for other contributions reaching 2.1 and 1.7 kJ·mol−1 for the S(VI) compounds SO3 and (CH3)2SO2, respectively. However, such a large deviation is not observed for H2SO4.
For the compounds with all contributions known, the standard deviation for the ΔEhoc deviations from additivity is greater than those of individual contributions considered above. This indicates that post-CCSD(T) and other higher-order corrections do not partially cancel one another. While the statistics is limited, the S(IV) compounds, SO2 and SSO, have the largest negative deviations close to −1.4 kJ·mol−1, the only S(VI) compound, SO3, has the largest positive deviation of 2.8 kJ·mol−1, and the S(II) compounds lie in-between.
Comparison of local and canonical CCSD(T) energies
The local and canonical CCSD(T)/aug-cc-pv(Q+d)Z energies for sulfur compounds are compared in Table 2. For flexible molecules exibiting multiple conformations, the computations were conducted for the lowest-energy conformer. In all cases except SO3 and H2SO4, the local energies are more negative than their canonical counterparts. If one uses the C, H, and O atomic contributions obtained earlier,18 the average difference between the additive and computed values is close to zero for the S(II) and S(IV) compounds. The S(VI) compounds have the deviations ranging from −(1.4 to 2.1) kJ·mol−1. For SO3, the deviation of −1.9 kJ·mol−1 is partially cancelled by a relatively large non-additivity of the ΔEhoc term. Although comprehensive statistics is not available, one would expect that the corrections for different atomic types due to contributions beyond frozen-core CCSD(T) or the use of local approximation will be small for organosulfur compounds.
Table 2:
Deviations of local CCSD(T) energies from canonical values
| compound | E(canonical) / Hartree | ΔEloc / (kJ·mol−1)a | dev. / (kJ·mol−1)b |
|---|---|---|---|
| hydrogen sulfide | −398.958965 | −0.3 | 0.2 |
| sulfur dioxide | −548.062961 | −0.5 | −0.1 |
| sulfur trioxide | −623.191638 | 0.9 | −1.9 |
| carbon oxide sulfide | −510.976509 | −1.0 | 0.2 |
| carbon disulfide | −833.561382 | −0.4 | 0.0 |
| disulfur monoxide | −870.653730 | −0.9 | 0.6 |
| sulfuric acid | −699.594130 | 0.7 | −2.1 |
| dimethyl sulfoxide | −552.588746 | −1.2 | −0.5 |
| dimethyl sulfone | −627.764576 | −0.6 | −1.4 |
| Standard deviation | 0.7 | 1.5 |
ΔEloc = E(local) – E(canonical).
Additive value minus the computed value. The additive values were found as18 kJ·mol−1 = −0.46nC – 0.07nH – 0.32nO
Finite basis set vs. complete basis set LCCSD(T) energies
CBS extrapolation of LCCSD(T) energies is potentially beneficial for the accuracy of the predicted ΔfH° values. In the protocol considered here, the QZ-quality basis set is used, and two two-point extrapolation schemes, TZ-QZ and QZ-5Z, are possible. The standard uncertainty of prediction with a TZ-quality basis set was 4.6 kJ·mol−1 as compared to 1.4 kJ·mol−1 with a QZ-based one.17 The uncertainty will propagate to the extrapolated energy thus making the scheme unacceptable for the present accuracy requirements. This leaves the QZ-5Z extrapolation as the only choice. It significantly increases the computational time, which is critical for a protocol designed to be used on a scale of thousands molecules. Therefore, the use of extrapolation should be limited to the cases where it is truly needed.
Initially, the extrapolation was applied to the reference set of 42 CHON molecules used to parameterize this protocol.18 The LCCSD(T)/CBS energies were then used to determine the effective enthalpies: h(C saturated or aromatic) = −99924.23 kJ·mol−1, h(C unsaturated) = −99923.22 kJ·mol−1, h(H) = −1525.10 kJ·mol−1, h(O) = −197170.63 kJ·mol−1, and h(N) = −143634.18 kJ·mol−1. The standard uncertainty of the prediction decreased insignificantly, from 1.25 kJ·mol−1 for the QZ version to 1.18 kJ·mol−1. Therefore, the use of extrapolation with this protocol for CHON-compounds is not justified.
Inorganic species are used for analysis of sulfur compounds (Table 3) because their experimental enthalpies of formation are deemed most reliable. Some challenges with the experimental data for organic compounds containing S(IV) and S(VI) are discussed below. The effective enthalpies h(S) determined with the QZ basis set vary from (−7.5 to 8.9) kJ·mol−1 from the average value of −1044341.3 kJ·mol−1. Two observations can be made based on these results. First, multiple corrections for different atomic types will be needed with this basis. Second, while Δh(SH) and Δh(S=) are expected to differ by about 2 kJ·mol−1, Δh(S(IV)) and Δh(S(VI)) are expected to be about (10 to 15) kJ·mol−1 more positive.
Table 3:
Effect of basis set extrapolation on the effective enthalpy of sulfur
| compound | Exp. / (kJ·mol−1) | Δ(QZ) / (kJ·mol−1)a | Δ(CBS) / (kJ·mol−1)b |
|---|---|---|---|
| hydrogen sulfide | −(20.6 ± 0.5) | −7.5 | −1.5 |
| sulfur dioxide | −(296.8 ± 0.2) | 3.3 | 1.9 |
| sulfur trioxide | −(395.9 ± 0.7) | 6.7 | 0.1 |
| carbon oxide sulfide | −(141.7 ± 2.0) | −5.2 | −0.4 |
| carbon disulfide | 116.7 ± 1.0 | −5.7 | −1.2 |
| disulfur monoxide | −(56.0 ± 1.4) | −0.4 | 1.7 |
| sulfuric acid | −(732.7 ± 2.0) | 8.9 | −0.5 |
deviation from the average value of h(S) = −1044341.3 kJ·mol−1;
deviation from the average value of h(S) = −1044370.3 kJ·mol−1
The deviations from the average value h(S) = −1044370.3 kJ·mol−1 for CBS energies are below ±2 kJ·mol−1 and do not reveal any structure-related patterns. These deviations are comparable to the experimental uncertainties. Therefore, a single parameter for sulfur may be sufficient in this case to predict the enthalpy of formation for organosulfur compounds within the target accuracy. This also implies that, based on presented, albeit rather limited, statistics, the basis set size appears to be a dominant factor for describing sulfur-containing compounds with chemical accuracy within an atom-equivalent additivity scheme. The same applies to calculation of reaction enthalpies, especially those involving compounds with different oxidation states of sulfur: accurate results would require the basis sets well above the quadruple-zeta quality, which would rule out many popular composite budget methods.
Thiols and sulfides
Most results for thiols and a significant portion of data for sulfides originate from the Bartesville laboratory. Some measurements were conducted at Lund, Belfast, and Madrid. We estimated the lower uncertainty limit of our predictions using values from Tables 4 and 5. Two preliminary parameters, h(S) = −1044348.6 kJ·mol−1 and Δh(SH) = −1.2 kJ·mol−1, were found by the least-squares fitting of experimental data with Eq. 1 and LCCSD(T)/aug-cc-pV(Q+d)Z energies. The standard uncertainty of the fit was 1.3 kJ·mol−1. Therefore, in the best-case scenario, the enthalpies of formation for the sulfur-containing substances can be predicted within ~2.6 kJ·mol−1 with a 0.95 probability. This estimate is comparable to the expanded uncertainty of about 3 kJ·mol−1 estimated for the enthalpies of formation of medium-sized CHON-compounds predicted with this method.18 This also provides the scale for accessing the quality of CHONS experimental data. If the experimental ΔfH° for a compound significantly deviates from the predicted value, uncertainty of the experimental value should be increased and the value itself should not be used for parameterization of the method.
Table 5:
Thermochemical properties of sulfides at T = 298 K
| Chemical Name | CASRN | Phase at T | Referencesa | |||
|---|---|---|---|---|---|---|
| Training set | ||||||
| 2-thiapropane (2.1) | 75-18-3 | l | −(37.4 ± 0.6) | 99,116; 48 | −0.4 | −1.5 |
| thiacyclobutane (2.2) | 287-27-4 | l | 61.0 ± 1.1 | 117,118; 119 | 1.0 | −0.1 |
| 2-thiabutane (2.3) | 624-89-5 | l | −(59.3 ± 1.1) | 100; 120 | −1.3 | −2.3 |
| thiomorpholine (2.4) | 123-90-0 | l | 8.6 ± 2.7 | 90 | 3.2 | 2.7 |
| S-ethyl ethanethioate (2.5) | 625-60-5 | l | −(227.8 ± 0.9) | 121; 122 | 1.9 | 1.0 |
| 2-thiapentane (2.6) | 3877-15-4 | l | −(81.9 ± 1.0) | 102; 108 | 0.0 | −0.6 |
| 3-methyl-2-thiabutane (2.7) | 1551-21-9 | l | −(90.2 ± 0.8) | 102; 123 | 0.5 | 0.2 |
| thiacyclohexane (2.8) | 1613-51-0 | l | −(63.3 ± 0.8) | 118,124; 124 | 0.2 | −0.5 |
| 2-methylthiacyclopentane (2.9) | 1795-09-1 | l | −(63.9 ± 0.8) | 56; 56,111 | 0.5 | 0.1 |
| cyclopentyl-1-thiaethane (2.10) | 7133-36-0 | l | −(64.3 ± 1.0) | 56; 56,111 | 0.9 | 0.7 |
| 2,3-dihydrobenzo[b]thiophene (2.11) | 4565-32-6 | cr | 108.5 ± 1.3 | 125 | 0.5 | −0.6 |
| 2,2,4,4-tetramethyl-3-thiapentane (2.12) | 107-47-1 | l | −(188.6 ± 0.7) | 56,126; 127: 128 | −1.4 | 0.6 |
| thianthrene (2.13) | 92-85-3 | cr | 288.5± 2.3 | 129; 77,130: 131–133 | 1.5 | −1.3 |
| diphenyl sulfide (2.14) | 139-66-2 | l | 229.3 ± 1.9 | 134: 135; 134,136 | 2.6 | 1.7 |
| dicyclohexyl sulfide (2.15) | 7133-46-2 | l | −(182.8 ± 1.4) | 137; 138 | −0.6 | 1.6 |
| 1,3-dithiacyclohexane (2.16) | 505-23-7 | cr | 6.2 ± 2.3 | 51; 54: 51 | −2.0 | −3.9 |
| Testing set | ||||||
| thiacyclopentane (2.17) | 110-01-0 | l | −(33.6 ± 1.1) | 117,139; 139,140 | 0.7 | −0.3 |
| 3-thiapentane (2.18) | 352-93-2 | l | −(83.4 ± 0.9) | 102,116; 48 | 0.4 | −0.3 |
| 3-methylthiacyclopentane (2.19) | 4740-00-5 | l | −(60.2 ± 0.8) | 56; 56,111 | −1.9 | −2.2 |
| 3,3-dimethyl-2-thiabutane (2.20) | 6163-64-0 | l | −(121.0 ± 0.8) | 141 | −0.6 | −0.2 |
| 2,4-dimethyl-3-thiapentane (2.21) | 625-80-9 | l | −(141.6 ± 0.9) | 56,126; 48 | 0.2 | 0.8 |
| 1-phenyl-1-thiaethane (2.22) | 100-68-5 | l | 97.6 ± 0.8 | 56: 128 | −0.5 | −1.8 |
| N-methylphenothiazine (2.23) | 1207-72-3 | cr | 271.3 ± 4.1 | 92 | 1.9 | 1.8 |
| 1,4-dithiacyclohexane (2.24) | 505-29-3 | cr | 1.8 ± 2.4 | 52; 54: 52 | −2.2 | −3.9 |
Sources of combustion or reaction energies used to derive the condensed-phase and sources of enthalpies of sublimation or vaporization are separated by semicolon. Unused sources (if any) are listed after colon for each field;
, kJ·mol−1
The results from Lund are in excellent agreement with those from Bartesville. The agreement with the results from Belfast is somewhat worse: for benzylthiol, methyl phenyl sulfide, and diphenyl sulfide, the differences are (5 to 8) kJ·mol−1. Considering problems with the results for oxygenated compounds from this laboratory described below, the results from Bartesville were favored in these cases.
The gas-phase enthalpies of formation of 1,3-51 and 1,4-dithiacyclohexanes52 and 1,3,5-trithiacyclohexane53 were reported by the Madrid team. The sublimation enthalpies of the dithiacyclohexanes derived from the temperature-dependent vapor pressures significantly deviate from those obtained from similar measurements by De Wit et al.54 Comparison with the ab initio results suggests that the latter is likely preferable. 1,3,5-trithiacyclohexane is prone to polymerization at the conditions of sublimation experiments,54 and partial polymerization in the combustion experiments cannot be fully ruled out. Consequently, the experimental ΔfH° value for this compound was excluded.
Disulfides and aromatic sulfur substances
Short-chain dialkyl disulfides were studied in Bartesville.55,56 For isomeric dibutyl disulfides, the data from Belfast are available.57 The measurements for di-tert-butyl derivative were conducted in both laboratories. In this work, the dimethyl, diethyl, and di-tert-butyl derivatives are considered. In addition, the enthalpy of formation of diphenyl disulfide was recentely revised by Ramos et al.58 This improved value is used here in the training set and two diaminodiphenyl derivatives from the same work were included in the test set.
Sulfur is a part of an aromatic ring in thiophenes, thiazoles, and isothiazoles. The ΔfH° data are available for the former two groups. To maximize the diversity of the aromatic groups, multiple thiophene and thiazole derivatives were included in the training set (Table 6). These results were mostly obtained in the Bartesville and Porto laboratories.
Table 6:
Thermochemical properties of disulfides and compounds with an aromatic sulfur atom at T = 298 K
| Chemical Name | CASRN | Phase at T | Referencesa | |||
|---|---|---|---|---|---|---|
| Training set | ||||||
| 2,3-dithiabutane (3.1) | 624-92-0 | l | −(23.8 ± 0.9) | 55,116; 48 | −1.5 | −2.5 |
| 3,4-dithiahexane (3.2) | 110-81-6 | l | −(74.2 ± 1.0) | 55,116; 116,142,143 | −2.4 | −2.7 |
| 2,2,5,5-tetramethyl-3,4-dithiahexane (3.3) | 110-06-5 | l | −(199.4 ± 1.2) | 56,144; 143 | −3.9 | −1.4 |
| diphenyl disulfide (3.4) | 882-33-7 | cr | 243.7 ± 3.1 | 58: 135,145; 58 | 2.1 | 3.2 |
| 2-aminothiazole (3.5) | 96-50-4 | cr | 144.9 ± 2.1 | 91 | −1.0 | −2.3 |
| 4-cyanothiazole (3.6) | 1452-15-9 | cr | 294.1 ± 1.0 | 146 | 2.6 | 3.3 |
| thiophene (3.7) | 110-02-1 | l | 115.4 ± 0.6 | 118,147; 48 | 1.0 | −0.4 |
| 4-methylthiazole (3.8) | 693-95-8 | l | 112.1 ± 0.8 | 146 | −0.5 | −1.4 |
| 2-methylthiophene (3.9) | 554-14-3 | l | 83.6 ± 0.9 | 148; 48 | 1.3 | 0.3 |
| benzothiazole (3.10) | 95-16-9 | l | 203.5 ± 1.0 | 76,149; 76,149,150 | −2.9 | −3.5 |
| 2,2’-dithiophene (3.11) | 492-97-7 | cr | 247.5 ± 2.7 | 151 | −1.7 | −3.1 |
| benzo[b]thiophene (3.12) | 95-15-8 | cr | 168.2 ± 1.0 | 56; 152: 153,154 | 1.5 | 0.8 |
| 2,5-dimethylbenzothiazole (3.13) | 95-26-1 | cr | 120.4 ± 2.5 | 76 | 2.6 | 2.7 |
| dibenzothiophene (3.14) | 132-65-0 | cr | 213.7 ± 1.5 | 56,155: 156 ; 134,155: 154 | −0.9 | −1.5 |
| Testing set | ||||||
| 2,2’-disulfanediyldianiline (3.15) | 1141-88-4 | cr | 226.0 ± 4.3 | 58 | 2.0 | 4.9 |
| 4,4’-disulfanediyldianiline (3.16) | 722-27-0 | cr | 240.5 ± 5.1 | 58 | 7.4 | 8.0 |
| 3-methylthiophene (3.17) | 616-44-4 | l | 82.9 ± 0.9 | 157; 48,157 | 1.7 | 0.8 |
| 2-benzothiazolamine (3.18) | 136-95-8 | cr | 179.8 ± 1.9 | 91: 95; 91 | 3.2 | 2.6 |
| 2-methylbenzothiazole (3.19) | 120-75-2 | l | 140.5 ± 2.8 | 76 | 14.3 | 14.1 |
| 5-fluoro-2-methylbenzothiazole (3.20) | 399-75-7 | l | −(47.0 ± 4.3) | 94 | 9.1 | |
| benzo[b]thiophene-2-carboxylic acid (3.21) | 6314-28-9 | cr | −(208.6 ± 3.8) | 96 | 4.0 | 3.5 |
| 1,2,3,4-tetrahydrodibenzothiophene (3.22) | 16587-33-0 | l | 102.0 ± 1.5 | 158 | −0.8 | −0.1 |
Sources of combustion or reaction energies used to derive the condensed-phase and sources of enthalpies of sublimation or vaporization are separated by semicolon. Unused sources (if any) are listed after colon for each field;
, kJ-mol−1
Substances with sulfur double bonds, S(IV), and S(VI)
Organic compounds containing S(IV) and S(VI) typically have S=O bonds, while a C=S bond is normally seen in the S(II) compounds. The compounds with a C=S bond have been actively investigated over the past five years.59–63 The S(IV) and S(VI) compounds for which the ΔfH° values are available include sulfoxides, sulfones, sulfites, sulfates, and sulfonamides. Dozens of compounds of this group have been studied. Unfortunately, the number of reliable ΔfH° values for the gas phase is surprisingly low. Those for dimethyl sulfoxide64,65 and thiacyclopentane 1,1-dioxide (sulfolane)66 included in the training set are the only reliable values that we are aware of. The enthalpy of vaporization reported for the latter was revised using the ideal-gas enthalpies computed in this work. Inconsistencies between the computed and experimental data for this group are discussed below.
Inorganic species
The enthalpies of formation for gaseous H2S and SO2 recommended by CODATA67 were used in this work. of SSO, SO3, CS2, COS, H2SO4 were taken from Ref. 68. The JANAF69,70 values for SO2, H2S, SSO, SO3, CS2, and H2SO4 are consistent with the recommendations above. However, the JANAF value for COS is 3.3 kJ·mol−1 less negative than the more recent one. is equal to the enthalpy of sublimation of rhombic sulfur. This value was derived using the temperature-dependent vapor pressures available in multiple works. The sublimation enthalpy of S8 and its uncertainty was evaluated in this work using the original data71,72 and ΔsublCP = −26 J·K−1·mol−1.73
Fehèr et al. published a series of papers on thermodynamic properties of hydrogen polysulfides. Out of these compounds, H2S2 is considered here. The enthalpy of its decomposition74 H2S2(l) = H2S(g) + 1/8 S8(rh) at atmospheric pressure was combined with the calorimetrically determined vaporization enthalpy75 and the CODATA value of to obtain . Some properties were reported at T = 293 K and adjusted to T = 298 K using the gas-phase heat capacities of S(rh) and H2S(g) recommended by CODATA and Cp(H2S2) calculated in this work.
Final regression and comparison with experimental data
During the parameter fitting procedure, equations for each compound were set on per sulfur atom basis. This was necessary to avoid the bias toward compounds with the number of sulfur atoms greater than one (e.g. S8). Initially, seven parameters described above were considered. Following the analysis, the Δh(SH) value was found to be statistically insignificant. The difference between Δh(S(arom)) and Δh(SH=) was very small, below 0.2 kJ·mol−1. The difference with was close to 0.3 kJ·mol−1, and also not statistically significant. Therefore, these three parameters were combined into one. Finally, the definitive estimation of Δh(S(IV)) and Δh(S(VI)) was challenging as the amount of reliable data for the S(IV) and S(VI) compounds is extremely limited. Form the initial analysis, it was found that Δh(S(VI)) is approximately twice as large as Δh(S(IV)) and, instead of having two independent parameters, the ratio was imposed.
The final least-squares fit of the experimental data with these three parameters resulted in the following values:
The computed values are compared with their experimental counterparts in Figures 2,3, and 4. Generally, the computed values are in a very good agreement with the experiments, with the standard deviation for the training set of 1.8 kJ·mol−1. For the testing set, the standard deviation is 4.8 kJ·mol−1, mainly because of five outliers in recent publications (Table 9). It comes close to that of the training set, 2.3 kJ·mol−1, if they are removed.
Figure 2:

Comparison of the computed (QZ version) and experimental enthalpies of formation for thiols (circles) and sulfides (diamonds). Filled and empty symbols are used for the training and testing set, respectively.
Figure 3:

Comparison of the computed (QZ version) and experimental enthalpies of formation for compounds with the sulfur double bond, S(IV), and S(VI) (circles); disulfides and aromatic compounds (diamonds); and inorganic compounds (triangles). Filled and empty symbols are used for the training and testing set, respectively.
Figure 4:

Comparison of the computed (QZ version) and experimental enthalpies of formation of compounds, for which the combustion energies were published after 2013: diamonds, determined in the Porto laboratory;59–61,63,76,90–94 circles, determined in the Puebla laboratory.58,62,95,96 The value for 1,3-dihydroimidazole-2-thione has a deviation of −32 kJ·mol−1 and is not shown
A single parameter, h(S) = −(1044371.95 ± 0.30) kJ·mol−1, was found to be statistically significant for the CBS version. The computed and experimental values are compared in Figures 5 and 6. The standard deviation of 2.2 kJ·mol−1 for the training set becomes 4.3 and 2.5 kJ·mol−1 for the testing set with the outliers and without them, respectively. Typically, the deviations between two versions is within 1.5 kJ per mole of sulfur atoms. Notable differences between the predictions are observed for disulfur monoxide (3.4 kJ·mol−1) and sulfur trioxide (3.0 kJ·mol−1). The reason of this behavior is unclear.
Figure 5:

Comparison of the computed (CBS version) and experimental enthalpies of formation for thiols (circles) and sulfides (diamonds). Filled and empty symbols are used for the training and testing set, respectively.
Figure 6:

Comparison of the computed (CBS version) and experimental enthalpies of formation for compounds with the sulfur double bond, S(IV), and S(VI) (circles); disulfides and aromatic compounds (diamonds); and inorganic compounds (triangles). Filled and empty symbols are used for the training and testing set, respectively.
The expanded uncertainty can be estimated as follows:
| (6) |
The first term in the right-hand side of Eq. 6 is a contribution associated with C, H, O, and N atoms and evaluated as described earlier.18 This contribution includes uncertainty of the model as well as uncertainties of the effective enthalpies of atoms. For the considered compounds, U(CHON) is close to (2.5 to 3.0) kJ·mol−1. U(h(S)) = 0.6 kJ·mol−1 for the CBS version of the protocol. While a rigorous evaluation is not possible, we also expect the uncertainty for the QZ version to be close to this value.
The results obtained with the QZ version of the protocol will be used in further analysis. The use of the CBS results leads to similar conclusions.
A relatively large deviation of 7.4 kJ·mol−1 is observed for 4,4’-disulfanediyldianiline while the deviation for the 2,2’- isomer is only 2.0 kJ·mol−1. The computed enthalpies of formation for two 1,3-dihydro-2H-benzimidazole-2-thione derivatives deviate from the experimental values reported by Perdomo et al.62 by about 6 kJ·mol−1. The ΔfH° value for the third compound from that work, 1-methyl-1,3-dihydro-2H-imidazole-2-thione, is consistent with the computations. The computed ΔfH° for 1,3-dihydro-2H-benzimidazole-2-thione itself is in good agreement with the experimental61 value reported by the Porto lab. Also, it was supported by the G3-theory computations in the original work.
Silva et al.76 reported the enthalpies of formation of three benzothiazole derivatives. A good agreement between the computed and experimental values is observed for benzothiazole and 2,5-dimethylbenzothiazole (Table 6). However, the unexpected difference of 14 kJ·mol−1 is seen for the 2-methyl derivative. Thus, we believe the computed value should be preferred for the latter. The deviation for 5-fluoro-2-methylbenzothiazole is about 9 kJ·mol−1. This compound is unusual for the combustion measurements because it has both fluorine and sulfur atoms. As such, some reference values required to adjust the experimental results to the standard state are not available in the literature. Thus, the comparison experiments had to be used to derive the ΔfH° value, likely causing an increase in uncertainty. Considering all these factors, the 9 kJ·mol−1 difference appears satisfactory.
The reason for a −6.0 kJ·mol−1 difference for phenoxathiin is unclear. Although the sublimation enthalpy of this compound have been measured in several works, its combustion energy has been reported only once.77 Additional measurements would be very helpful.
Based on their own G3-theory calculations, the authors of the original publication on 1,3-dihydroimidazole-2-thione63 have also concluded that their value is about 30 kJ·mol−1 higher than the ab initio result. The inconsistency was explained by suggesting formation of the thiol tautomer whose ΔfH° was in excellent agreement with the calculations. However, the combustion experiments were conducted with the crystalline sample, and the crystal phase was confirmed to contain the thione form.63 Formation of the tautomer with the Gibbs energy of about 30 kJ·mol−1 in the sublimation experiments does not appear plausible, and, again, new experimental data on the gas-phase composition of this compound would be beneficial to address the existing inconsistency.
Further, the problems with the enthalpies of formation of the S(IV) and S(VI) compounds are discussed. For dimethyl sulfone (methylsulfonyl methane), two consistent values for the solid-phase ΔfH° have been published.64,78 The enthalpy of sublimation needed to convert this solid-phase value to the gas-phase ΔfH° was derived in this work as follows. Vapor pressure over the liquid sulfone was measured using a Bourdon gauge as a null manometer in the temperature range of (385 to 523) K and reported in the form of the Antoine equation. 79 If one neglects non-ideality of the gas phase, the enthalpy of vaporization is estimated from vapor pressure slope as ΔvamΗHm(454 K) = (54.2 ± 1.3) kJ·mol−1. The enthalpy difference Hm(l, 454 K) − Hm(cr, 298 K) = 43.2 kJ·mol−1 can be determined using the linearly extrapolated calorimetric data of Clever and Westrum.80 The gas-phase enthalpy change, Hm(g, 454 K) − Hm(g, 298 K) = 18.2 kJ·mol−1 was calculated using the present theoretical model described above. The final enthalpy of sublimation obtained in this manner is ΔsublHm(298K) = (79.2 ± 1.4) kJ·mol−1. The resulting estimated uncertainty is fairly low; yet, a large difference between the computed and experimental value exists. It is most likely caused by ΔsublHm value and the issue can only be resolved by performing new measurements, corroborating or rejecting the available limited data.79 The temperature-dependent vapor pressures of this compound were recently estimated using thermogravimetric analysis and reported as coefficients of the Antoine equation.81 An unphysical value of +958 K reported for the C coefficient indicates that these results are in error.
The enthalpies of formation of other aliphatic and aromatic sulfones were determined in the Belfast lab. Some original data for sulfones have never been published and are only available in the review of Cox and Pilcher.46 Di-n-alkyl sulfones are considered here as an example (Figure 7). The experimental increment of the gas-phase ΔfH° per CH2 group varies from −(8 to 36) kJ·mol−1 between different homologues. According to the computations in this work, it is close to −24 kJ·mol−1 for the short-chain compounds (n(C) < 5). For the longer alkyl chains, the increment similar to that of n-alkanes (about 21 kJ·mol−1) is expected. No notable systematic shift between the computed and experimental values is observed (Figure 7). However, significant data scatter indicates that the expanded uncertainty of the experimental data should be close to 10 kJ·mol−1.
Figure 7:

Deviation of the estimated enthalpies of formation of normal dialkyl sulfones (blue circles), sulfites (yellow squares), and sulfates (red diamonds) from their experimental counterparts as a function of carbon atom number. The experimental values obtained in the Belfast lab are taken from the compilation of Cox and Pilcher.46 The original data were partially published in Refs. 78, 79, 82, and 97. The repeatability-based expanded uncertainties stated by the authors are shown. The enthalpies of formation at n < 5 and for n-propyl sulfate were computed using the QZ version of the protocol described in this work. The values at other n were estimated using a −21 kJ·mol−1 increment per CH2 group.
The only work on sulfites and sulfates was published by Mackle and Steele.82 The experimental increment per CH2 group varies from −(28 to 41) kJ·mol−1 for the short-chain homologues compared to −34 and −35 kJ·mol−1 computed for the sulfites and sulfates, respectively. Because of significant data scatter, it cannot be concluded whether the systematic deviations between the computed and experimental values are due to incorrect h(S) or large experimental uncertainties. Consequently, the data on sulfones, sulfites, and sulfates were excluded from further consideration. This leaves sulfuric acid as the only compound with both S=O and S-O bonds that has a reliable value of ΔfH°. Generation of new experimental data for organic sulfites and sulfates is critical to closing this neglected gap in thermochemistry of sulfur and would contribute greatly to the development and validation of theoretical methods.
The energy of combustion of another candidate, benzenesulfonamide, has been reported in two works.83,84 With the use the enthalpy of sublimation from Ref. 83, the values of the resulting gas-phase enthalpies of formation from two sources differ by 23 kJ·mol−1. Comparison with the prediction of the current model suggests that the result of Matos et al.83 is more favorable (−5 kJ·mol−1 deviation), while the value of Flores et al.84 shows the deviation of 18 kJ·mol−1.
Thiacyclohexane 1,1-dioxide85 has a −9.8 kJ·mol−1 deviation from the experimental value. The computed enthalpies of formation for similar compounds have a −9 kJ·mol−1 average deviation from the experimental values86–89 originating from the same laboratory. The origins of this behavior are unclear and also need to be resolved by conducting new measurements.
From the major data contributors discussed above, only the laboratories in Porto and Puebla are still active. The computed ΔfH° values are compared with experimental data published by these two groups after 2013 (Figure 4). For three values from the Porto lab and one value (benzenesulfonamide) from Puebla, the deviations exceed +9 kJ·mol−1 (all are discussed above). The other experimental results are consistent with the expected expanded uncertainty of the predicted values, 3 kJ·mol−1. The data from Porto are scattered around zero with the average deviation of 0.6 kJ·mol−1. The results from the Puebla lab are systematically more negative than the predictions with the average deviation of 4.1 kJ·mol−1.
Summary
The extension of our previously developed LCCSD(T)-based protocol18 for prediction of the gas-phase ΔfH° for closed-shell compounds to sulfur-containing compounds is presented. It shares the computational efficiency of the original formulation, thus allowing rapid evaluation and large-scale processing applications. The accuracy and generalization of the method is hindered by (somewhat unexpectedly) limited amount of reliable data for the diverse classes of sulfur-containing organic compounds. Numerous, from obvious to suspected, experimental issues have been identified. Furthermore, surprising gaps in data coverage for compound classes of both practical and scientific interest (e.g., sulfites and sulfates) were found. This causes problems in model parametrization as well as in its proper validation and accuracy assessment.
Formal comparison with a large set of available data critically-evaluated in this work yields the standard deviation of about 2 kJ·mol−1, suggesting a 95% confidence uncertainty of about 4 kJ·mol−1. This value, however, represents an upper estimate. The actual uncertainty, considering the scatter in the experimental data, is likely to be closer to 3 kJ·mol−1, as evidenced by the analysis of the data subsets exhibiting lower levels of experimental uncertainties.
Table 7:
Thermochemical properties of organic compounds with the sulfur double bond, S(IV), and S(VI) at T = 298 K
| Chemical Name | CASRN | Phase at T | Referencesa | |||
|---|---|---|---|---|---|---|
| Training set | ||||||
| thioacetamide (4.1) | 62-55-5 | cr | 12.9 ± 1.1 | 159: 160; 159,160 | 1.2 | 2.0 |
| imidazolidine-2-thione (4.2) | 96-45-7 | cr | 71.0 ± 2.1 | 59 | −3.3 | −3.1 |
| 1-aza-2-cycloheptanethione (4.3) | 7203-96-5 | cr | −(3.5 ± 2.5) | 60 | 2.1 | 4.0 |
| 1,3-dihydro-2H-benzimidazole-2-thione (4.4) | 583-39-1 | cr | 175.1 ± 2.2 | 61: 161; 61 | −2.8 | −3.0 |
| 1-methyl-1,3-dihydro-2H-imidazole-2-thione (4.5) | 60-56-0 | cr | 116.3 ± 2.3 | 62 | 1.8 | 1.6 |
| dimethyl sulfoxide (4.6) | 67-68-5 | l | −(150.6 ± 1.0) | 64,65; 162 | −3.1 | −2.5 |
| thiacyclopentane 1,1-dioxide (4.7) | 126-33-0 | lc | −(358.3 ± 1.8) | 66; 66,162,163 | −1.0 | 1.0 |
| Testing set | ||||||
| 1,3-dihydro-5-methoxy-2H-benzimidazole-2-thione (4.8) | 37052-78-1 | cr | 19.6 ± 2.8 | 62 | 5.2 | 5.9 |
| 3H-1,3-benzothiazole-2-thione (4.9) | 149-30-4 | cr | 205.5 ± 3.3 | 164: 161; 164 | 3.4 | 2.0 |
| 5-amino-1,3-dihydro-2H-benzimidazole-2-thione (4.10) | 2818-66-8 | cr | 174.6 ± 3.7 | 62 | 6.6 | 6.1 |
| 2-thiobarbituric acid (4.11) | 504-17-6 | cr | −(278.2 ± 4.5) | 165d, 93; 165: 166 | 2.5 | 3.5 |
Sources of combustion or reaction energies used to derive the condensed-phase and sources of enthalpies of sublimation or vaporization are separated by semicolon. Unused sources (if any) are listed after colon for each field;
, kJ·mol−1;
supercooled;
revised using the NIST Combustion Calorimetry Tool47
Table 8:
Thermochemical properties of S-containing inorganic compounds at T = 298 K
| Chemical Name | CASRN | Phase at T | Referencesa | |||
|---|---|---|---|---|---|---|
| hydrogen sulfide (5.1) | 7783-06-4 | g | −(20.6 ± 0.5) | 67 | −0.1 | 0.2 |
| sulfur dioxide (5.2) | 7446-09-5 | g | −(296.8 ± 0.2) | 67 | 2.2 | 3.6 |
| sulfur trioxide (5.3) | 7446-11-9 | l | −(395.9 ± 0.7) | 68 | −1.2 | 1.8 |
| sulfuric acid (5.4) | 7664-93-9 | l | −(732.7 ± 2.0) | 68 | 2.7 | 1.2 |
| hydrogen disulfide (5.5) | 13465-07-1 | l | 15.6 ± 1.4 | 74; 75: 167 | 0.6 | 0.8 |
| disulfur monoxide (5.6) | 20901-21-7 | l | −(56.0 ± 1.4) | 68 | 0.3 | 6.8 |
| octasulfur (5.7) | 10544-50-0 | cr | 99.9 ± 0.5 | 71,72 | −2.6 | −4.7 |
| carbon oxide sulfide (5.8) | 463-58-1 | g | −(141.7 ± 2.0) | 68 | 0.4 | 1.2 |
| carbon disulfide (5.9) | 75-15-0 | l | 116.7 ± 1.0 | 68 | −0.2 | 1.0 |
Sources of combustion or reaction energies used to derive the condensed-phase and sources of enthalpies of sublimation or vaporization are separated by semicolon. Unused sources (if any) are listed after colon for each field;
, kJ-mol−1
Acknowledgement
Trade names are provided only to specify procedures adequately and do not imply endorsement by the National Institute of Standards and Technology. Similar products by other manufacturers may be found to work as well or better. The authors declare no competing financial interest.
References
- (1).Pedley JB Thermochemical Data and Structures of Organic Compounds; TRC Data Series; Thermodynamics Research Center, College Station, TX, 1994. [Google Scholar]
- (2).Gertig C; Leonhard K; Bardow A Computer-aided molecular and processes design based on quantum chemistry: current status and future prospects. Curr. Opin. Chem. Eng 2020, 27, 89–97. [Google Scholar]
- (3).Diky V; Bazyleva A; Paulechka E; Magee JW; Martinez V; Riccardi D; Kroenlein K Validation of thermophysical data for scientific and engineering applications. J. Chem. Thermodyn 2019, 133, 208–222. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (4).Irikura KK Extracting Thermochemical Information from ab initio Data. In Energetics of Stable Molecules and Reactive Intermediates; Minas da Piedade ME, Ed.; Kluwer Academic Publishers: Dordrecht, the Netherlands, 1999; pp 353–372. [Google Scholar]
- (5).Hampel C; Werner H-J Local treatment of electron correlation in coupled cluster theory. J. Chem. Phys 1996, 104 , 6286–6297. [Google Scholar]
- (6).Schütz M; Werner H-J Local perturbative triples correction (T) with linear cost scaling. Chem. Phys. Lett 2000, 318, 370–378. [Google Scholar]
- (7).Riplinger C; Neese F An efficient and near linear scaling pair natural orbital based local coupled cluster method. J. Chem. Phys 2013, 138, 034106. [DOI] [PubMed] [Google Scholar]
- (8).Liakos DG; Neese F Is It Possible To Obtain Coupled Cluster Quality Energies at near Density Functional Theory Cost? Domain-Based Local Pair Natural Orbital Coupled Cluster vs Modern Density Functional Theory. J. Chem. Theory Comput 2015, 11, 4054–4063. [DOI] [PubMed] [Google Scholar]
- (9).Liakos DG; Sparta M; Kesharwani MK; Martin JML; Neese F Exploring the Accuracy Limits of Local Pair Natural Orbital Coupled-Cluster Theory. J. Chem. Theory Comput 2015, 11, 1525–1539. [DOI] [PubMed] [Google Scholar]
- (10).Riplinger C; Pinski P; Becker U; Valeev EF; Neese F Sparse maps—A systematic infrastructure for reduced-scaling electronic structure methods. II. Linear scaling domain based pair natural orbital coupled cluster theory. J. Chem. Phys 2016, 144, 024109. [DOI] [PubMed] [Google Scholar]
- (11).Nagy PR; Samu G; Kállay M An Integral-Direct Linear-Scaling Second-Order Møller–Plesset Approach. J. Chem. Theory Comput 2016, 12, 4897–4914. [DOI] [PubMed] [Google Scholar]
- (12).Nagy PR; Kállay M Optimization of the linear-scaling local natural orbital CCSD(T) method: Redundancy-free triples correction using Laplace transform. J. Chem. Phys 2017, 146, 214106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (13).Nagy PR; Samu G; Kállay M Optimization of the Linear-Scaling Local Natural Orbital CCSD(T) Method: Improved Algorithm and Benchmark Applications. J. Chem. Theory Comput 2018, 14, 4193–4215. [DOI] [PubMed] [Google Scholar]
- (14).van Alsenoy C Ab initio calculations on large molecules: The multiplicative integral approximation. J. Comput. Chem 1988, 9, 620–626. [Google Scholar]
- (15).Feyereisen M; Fitzgerald G; Komornicki A Use of approximate integrals in ab initio theory. An application in MP2 energy calculations. Chem. Phys. Lett 1993, 208, 359–363. [Google Scholar]
- (16).Weigend F; Häser M RI-MP2: first derivatives and global consistency. Theor. Chem. Acc 1997, 97, 331–340. [Google Scholar]
- (17).Paulechka E; Kazakov A Efficient DLPNO–CCSD(T)-Based Estimation of Formation Enthalpies for C-, H-, O-, and N-Containing Closed-Shell Compounds Validated Against Critically Evaluated Experimental Data. J. Phys. Chem. A 2017, 121, 4379–4387. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (18).Paulechka E; Kazakov A Efficient Estimation of Formation Enthalpies for Closed-Shell Organic Compounds with Local Coupled-Cluster Methods. J. Chem. Theory Comput 2018, 14, 5920–5932. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (19).Paulechka E; Kazakov A Critical Evaluation of the Enthalpies of Formation for Fluorinated Compounds Using Experimental Data and High-Level Ab Initio Calculations. J. Chem. Eng. Data 2019, 64, 4863–4875. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (20).Klippenstein SJ; Harding LB; Ruscic B Ab Initio Computations and Active Thermochemical Tables Hand in Hand: Heats of Formation of Core Combustion Species. J. Phys. Chem. A 2017, 121, 6580–6602. [DOI] [PubMed] [Google Scholar]
- (21).Mole SJ; Zhou X; Liu R Density Functional Theory (DFT) Study of Enthalpy of Formation. 1. Consistency of DFT Energies and Atom Equivalents for Converting DFT Energies into Enthalpies of Formation. J. Phys. Chem 1996, 100, 14665–14671. [Google Scholar]
- (22).Tirado-Rives J; Jorgensen WL Performance of B3LYP Density Functional Methods for a Large Set of Organic Molecules. J. Chem. Theory Comput 2008, 4, 297–306. [DOI] [PubMed] [Google Scholar]
- (23).Raghavachari K; Stefanov BB; Curtiss LA Accurate thermochemistry for larger molecules: Gaussian-2 theory with bond separation energies. J. Chem. Phys 1997, 106, 6764–6767. [Google Scholar]
- (24).Rossini FD, Ed. Experimental Thermochemistry; Interscience Publishers: New York, 1956; Vol. 1. [Google Scholar]
- (25).Kwart H; King K d-Orbitals in the Chemistry of Silicon, Phosphorus and Sulfur; Springer, Heidelberg, 1977. [Google Scholar]
- (26).Karton A; Rabinovich E; Martin JML; Ruscic B W4 theory for computational thermochemistry: In pursuit of confident sub-kJ/mol predictions. J. Chem. Phys 2006, 125, 144108. [DOI] [PubMed] [Google Scholar]
- (27).Feller D; Peterson KA; Dixon DA A survey of factors contributing to accurate theoretical predictions of atomization energies and molecular structures. J. Chem. Phys 2008, 129, 204105. [DOI] [PubMed] [Google Scholar]
- (28).Hahn DK; RaghuVeer KS; Ortiz JV Computational Tests of Quantum Chemical Models for Structures, Vibrational Frequencies, and Heats of Formation of Molecules with Phosphorus and Sulfur Atoms. J. Phys. Chem. A 2010, 114, 8142–8155. [DOI] [PubMed] [Google Scholar]
- (29).Dunning TH; Peterson KA; Wilson AK Gaussian basis sets for use in correlated molecular calculations. X. The atoms aluminum through argon revisited. J. Chem. Phys 2001, 114, 9244–9253. [Google Scholar]
- (30).Martin JML Heat of atomization of sulfur trioxide, SO3: a benchmark for computational thermochemistry. Chem. Phys. Lett 1999. , 310, 271–276. [Google Scholar]
- (31).Bell RD; Wilson AK SO3 revisited: Impact of tight d augmented correlation consistent basis sets on atomization energy and structure. Chem. Phys. Lett 2004, 394, 105–109. [Google Scholar]
- (32).Frisch MJ; Trucks GW; Schlegel HB; Scuseria GE; Robb MA; Cheeseman JR; Scalmani G; Barone V; Petersson GA; Nakatsuji H et al. Gaussian 16 Revision B.01 2016; Gaussian Inc., Wallingford, CT. [Google Scholar]
- (33).Parrish RM; Burns LA; Smith DGA; Simmonett AC; DePrince AE; Hohenstein EG; Bozkaya U; Sokolov AY; Remigio RD; Richard RM et al. Psi4 1.1: An Open-Source Electronic Structure Program Emphasizing Automation, Advanced Libraries, and Interoperability. J. Chem. Theory Comput 2017, 13, 3185–3197. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (34).Smith DGA; Burns LA; Simmonett AC; Parrish RM; Schieber MC; Galvelis R; Kraus P; Kruse H; Remigio RD; Alenaizan A et al. Psi4 1.4: Open-source software for high-throughput quantum chemistry. J. Chem. Phys 2020, 152, 184108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (35).Kállay M; Nagy PR; Mester D; Rolik Z; Samu G; Csontos J; Csóka J; Szabó PB; Gyevi-Nagy L; Hégely B et al. The MRCC program system: Accurate quantum chemistry from water to proteins. J. Chem. Phys 2020, 152, 074107. [DOI] [PubMed] [Google Scholar]
- (36).Kállay M; Rolik Z; Csontos J; Nagy P; Samu G; Mester D; Csóka J; Szabó B; Ladjánszki I; Szegedy L et al. MRCC, a quantum chemical program suite. http://www.mrcc.hu, 2019.
- (37).Karton A; Martin JML Comment on: “Estimating the Hartree—Fock Limit from Finite Basis Set Calculations” [Jensen F (2005) Theor Chem Acc 113:267]. Theor. Chem. Acc 2006, 115, 330–333. [Google Scholar]
- (38).Jensen F Estimating the Hartree–Fock Limit from Finite Basis Set Calculations. Theor. Chem. Acc 2005, 113, 267–273. [Google Scholar]
- (39).Halkier A; Helgaker T; Jørgensen P; Klopper W; Koch H; Olsen J; Wilson AK Basis-Set Convergence in Correlated Calculations on Ne, N2, and H2O. Chem. Phys. Lett 1998, 286, 243–252. [Google Scholar]
- (40).Douglas M; Kroll NM Quantum Electrodynamical Corrections to the Fine Structure of Helium. Ann. Phys 1974, 82, 89–155. [Google Scholar]
- (41).Hess BA Relativistic Electronic-Structure Calculations Employing a Two-Component No-Pair Formalism with External-Field Projection Operators. Phys. Rev. A 1986, 33, 3742–3748. [DOI] [PubMed] [Google Scholar]
- (42).Ye H; Mendolicchio M; Kruse H; Puzzarini C; Biczysko M; Barone V The challenging equilibrium structure of HSSH: another success of the rotational spectroscopy/quantum chemistry synergism. J. Mol. Struct 2020, 1211, 127933. [Google Scholar]
- (43).Wang H; Frenklach M Enthalpies of formation of benzenoid aromatic molecules and radicals. J. Phys. Chem 1993, 97, 3867–3874. [Google Scholar]
- (44).Melius CF Thermochemical Modeling: I. Application to Decomposition of Energetic Materials. In Chemistry and Physics of Energetic Materials; Bulusu SN, Ed.; Springer Netherlands: Dordrecht, 1990; pp 21–49. [Google Scholar]
- (45).Anantharaman B; Melius CF Bond Additivity Corrections for G3B3 and G3MP2B3 Quantum Chemistry Methods. J. Phys. Chem. A 2005, 109, 1734–1747. [DOI] [PubMed] [Google Scholar]
- (46).Cox JD; Pilcher G Thermochemistry of Organic and Organometallic Compounds; Academic Press: New York, 1970. [Google Scholar]
- (47).Paulechka E; Riccardi D; Bazyleva A; Ribeiro da Silva MDMC; Zaitsau D Corrections to Standard State in Combustion Calorimetry: An Update and a Web-Based Tool. J. Chem. Thermodyn 2021, 158, 106425. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (48).NIST ThermoData Engine Version 10.1 - Pure Compounds, Binary Mixtures and Reactions. NIST Standard Reference Database 103b, Gaithersburg, MD, 2020. https://www.nist.gov/mml/acmd/trc/thermodata-engine/srd-nist-tde-103b, accessed: January 30, 2021. [Google Scholar]
- (49).Karton A; Tarnopolsky A; Lamère J-F; Schatz GC; Martin JML Highly Accurate First-Principles Benchmark Data Sets for the Parametrization and Validation of Density Functional and Other Approximate Methods. Derivation of a Robust, Generally Applicable, Double-Hybrid Functional for Thermochemistry and Thermochemical Kinetics. J. Phys. Chem. A 2008, 112, 12868–12886. [DOI] [PubMed] [Google Scholar]
- (50).Alexeev Y; Windus TL; Zhan C; Dixon DA Accurate Heats of Formation and Acidities for H3PO4, H2SO4, and H2CO3 from ab initio Electronic Structure Calculations. Int. J. Quantum Chem 2005, 102, 775–784. [Google Scholar]
- (51).Roux MV; Dávalos JZ; Jiménez P; Flores H; Saiz J-L; Abboud J-LM; Juaristi E Structural Effects on the Thermochemical Properties of Sulfur Compounds: I. Enthalpy of Combustion, Vapour Pressures, Enthalpy of Sublimation, and Standard Molar Enthalpy of Formation in the Gaseous Phase of 1,3-Dithiane. J. Chem. Thermodyn 1999, 31, 635–646. [Google Scholar]
- (52).Dávalos JZ; Flores H; Jiménez P; Notario R; Roux MV; Juaristi E; Hosmane RS; Liebman JF Calorimetric, Computational (G2(MP2) and G3) and Conceptual Study of the Energetics of the Isomeric 1,3- and 1,4-Dithianes. J. Org. Chem 1999, 64, 9328–9336. [Google Scholar]
- (53).Roux MV; Jiménez P; Dávalos JZ; Notario R; Juaristi E Calorimetric and Computational Study of 1,3,5-Trithiane. J. Org. Chem 2001, 66, 5343–5351. [DOI] [PubMed] [Google Scholar]
- (54).De Wit HGM; Van Miltenburg JC; De Kruif CG Thermodynamic Properties of Molecular Organic Crystals Containing Nitrogen, Oxygen, and Sulphur 1. Vapour Pressures and Enthalpies of Sublimation. J. Chem. Thermodyn 1983, 15, 651–663. [Google Scholar]
- (55).Hubbard WN; Douslin DR; McCullough JP; Scott DW; Todd SS; Messerly JF; Hossenlopp IA; George A; Waddington G 2,3-Dithiabutane, 3,4-Dithiahexane and 4,5-Dithiaoctane: Chemical Thermodynamic Properties from 0 to 1000°K. J. Am. Chem. Soc 1958, 80, 3547–3554. [Google Scholar]
- (56).Good WD Enthalpies of Combustion of 18 Organic Sulfur Compounds Related to Petroleum. J. Chem. Eng. Data 1972, 17, 158–162. [Google Scholar]
- (57).Mackle H; McClean RTB Studies in the Thermochemistry of Organic Sulphides. Part 5.–Dithia-alkanes. Trans. Faraday Soc 1964, 60, 669–672. [Google Scholar]
- (58).Ramos F; Flores H; Hernández-Pérez JM; Sandoval-Lira J; Camarillo EA The Intramolecular Hydrogen Bond N—HS…in 2,2’-Diaminodiphenyl Disulfide: Experimental and Computational Thermochemistry. J. Phys. Chem. A 2017, 122, 239–248. [DOI] [PubMed] [Google Scholar]
- (59).Silva ALR; Ribeiro da Silva MDMC Comprehensive Thermochemical Study of Cyclic Five- and Six-Membered N,N’-Thioureas. J. Chem. Eng. Data 2017, 62, 2584–2591. [Google Scholar]
- (60).Freitas VLS; Ribeiro da Silva DMC Thermodynamic Properties of ϵ-Caprolactam and ϵ-Caprothiolactam. J. Chem. Thermodyn 2019, 132, 451–460. [Google Scholar]
- (61).Silva ALR; Ribeiro da Silva MDMC Energetic, Structural and Tautomeric Analysis of 2-Mercaptobenzimidazole. J. Therm. Anal. Calorim 2017, 129, 1679–1688. [Google Scholar]
- (62).Perdomo G; Flores H; Ramos F; Notario R; Freitas VLS; Ribeiro da Silva MDMC; Camarillo EA; Dávalos JZ Thermochemistry of R-SH Group in Gaseous Phase: Experimental and Theoretical Studies of Three Sulfur Imidazole Derivatives. J. Chem. Thermodyn 2018, 122, 65–72. [Google Scholar]
- (63).Silva ALR; Morais VMF; Ribeiro da Silva MDMC; Simões RG; Bernardes CES; Piedade MFM; Minas de Piedade ME Structural and Energetic Characterization of Anhydrous and Hemihydrated 2-Mercaptoimidazole: Calorimetric, X-Ray Diffraction, and Computational Studies. J. Chem. Thermodyn 2016, 95, 35–48. [Google Scholar]
- (64).Douglas TB Heats of Formation of Liquid Methyl Sulfoxide and Crystalline Methyl Sulfone at 18°. J. Am. Chem. Soc 1946, 68, 1072–1076. [Google Scholar]
- (65).Masuda N; Nagano Y; Sakiyama M Standard Molar Enthalpy of Formation of (CH3)2SO, Dimethylsulfoxide, by Combustion Calorimetry. J. Chem. Thermodyn 1994. , 26, 971–975. [Google Scholar]
- (66).Morais VMF; Matos MAR; Miranda MS; Liebman JF Surprises with Strain Energy and Sulpholane (Tetrahydrothiophene 1,1-Dioxide): A Combined Experimental and Theoretical Investigation. Mol. Phys 2004, 102, 525–530. [Google Scholar]
- (67).Cox JD; Wagman DD; Medvedev VA CODATA Key Values for Thermodynamics; Hemisphere Publishing Corp.: New York, 1989. [Google Scholar]
- (68).Gurvich LV; Veits IV; Medvedev VA; Khachkuruzov GA; Yungman VS; Bergman GA; Baibuz VF; Iorish VS; Yurkov GN; Gobov SI et al. Thermodynamic Properties of Individual Substances; Nauka: Moscow, 1978. [Google Scholar]
- (69).NIST-JANAF Thermochemical Tables. NIST Standard Reference Database 13, Gaithersburg, MD, 1998. https://janaf.nist.gov, accessed: January 30, 2021. [Google Scholar]
- (70).Dorofeeva OV; Iorish VS; Novikov VP; Neumann DB NIST-JANAF Thermochemical Tables. II. Three Molecules Related to Atmospheric Chemistry: HNO3, H2SO4, and H2O2. J. Phys. Chem. Ref. Data 2003, 32, 879–901. [Google Scholar]
- (71).Bradley RS Rates of Evaporation. IV. The Rate of Evaporation and Vapour Pressure of Rhombic Sulphur. Proc. Roy. Soc. A 1951, 205, 553–563. [Google Scholar]
- (72).Bricke C; Hartshorne NH; Stranks DR The Vapour Pressures and Latent Heats of Sublimation of α-, β-, and γ-Sulfur. J. Chem. Soc 1960, 1200–1209. [Google Scholar]
- (73).Thermal Constants of Substances. http://www.chem.msu.ru/cgi-bin/tkv.pl, accessed: January 30, 2021.
- (74).Fehér F; Winkhaus G Beiträge zur Chemie des Schwefels. 40. Zur Thermochemie der Sulfane: Bildungsenthalpien und Bindungsenergien. Z. Anorg. Allg. Chem 1957, 292, 210–223. [Google Scholar]
- (75).Fehér F; Hitzemann G Beiträge zur Chemie des Schwefels. 44. Verdampfungsenthalpien, Dampfdrucke, Siedepunkte, kritische Temperaturen und Drucke sowie Troutonsche Konstanten von Sulfanen. Z. Anorg. Allg. Chem 1958, 294, 50–62. [Google Scholar]
- (76).Silva ALR; Cimas A; Ribeiro da Silva MDMC Energetic Study of Benzothiazole and Two Methylbenzothiazole Derivatives: Calorimetric and Computational Approaches. J. Chem. Thermodyn 2014, 73, 3–11. [Google Scholar]
- (77).Steele WV; Chirico RD; Knipmeyer SE; Nguyen A The Thermodynamic Properties of Thianthrene and Phenoxathiin. J. Chem. Thermodyn 1993, 25, 965–992. [Google Scholar]
- (78).Busfield WK; Mackle H; O’Hare PAG Studies in the Thermochemistry of Sulphones. Part 2. – The Standard Heats of Formation of Sulphones of the Type RSO2CH3. Trans. Faraday Soc 1961, 57, 1054–1057. [Google Scholar]
- (79).Busfield WK; Ivin KJ; Mackle H; O’Hare PAG Studies in the Thermochemistry of Sulphones. Part 3. – Fusion and Vaporization Heats of Sulphones of the Type RSO2CH3. Trans. Faraday Soc 1961, 57, 1058–1063. [Google Scholar]
- (80).Clever HL; Westrum EF Dimethyl Sulfoxide and Dimethyl Sulfone. Heat Capacities, Enthalpies of Fusion, and Thermodynamic Properties. J. Phys. Chem 1970, 74, 1309–1317. [Google Scholar]
- (81).Weh R; de Klerk A Thermochemistry of Sulfones Relevant to Oxidative Desulfurization. Energy Fuels 2017, 31, 6607–6614. [Google Scholar]
- (82).Mackle H; Steele WV Studies in the Thermochemistry of Organic Sulphites and Sulphates. Part 1. – Heats of Combustion and Formation of Some Dialkyl Sulphites and Sulphates. Trans. Faraday Soc 1969, 65, 2053–2059. [Google Scholar]
- (83).Matos MAR; Miranda MS; Morais VMF; Liebman JF Saccharin: A Combined Experimental and Computational Thermochemical Investigation of a Sweetener and Sulfonamide. Mol. Phys 2005, 103, 221–228. [Google Scholar]
- (84).Flores H; Camarillo EA; Amador P Enthalpies of Combustion and Formation of Benzenesulfonamide and Some of Its Derivatives. J. Chem. Thermodyn 2012, 47, 408–411. [Google Scholar]
- (85).Roux MV; Temprado M; Jiménez P; Dávalos JZ; Notario R; Guzmán-Mejía R; Juaristi E Calorimetric and Computational Study of Thiacyclohexane 1-Oxide and Thiacyclohexane 1,1-Dioxide (Thiane Sulfoxide and Thiane Sulfone). Enthalpies of Formation and the Energy of the S=O Bond. J. Org Chem 2003, 68, 1762–1770. [DOI] [PubMed] [Google Scholar]
- (86).Roux MV; Temprado M; Jiménez P; Dávalos JZ; Notario R; Martín-Valcárel G; Garrido L; Guzmán-Mejía R; Juaristi E Thermochemistry of 1,3-Dithiacyclohexane 1-Oxide (1,3-Dithiane Sulfoxide): Calorimetric and Computational Study. J. Org. Chem 2004, 69, 5454–5459. [DOI] [PubMed] [Google Scholar]
- (87).Roux MV; Temprado M; Jiménez P; Notario R; Guzmán-Mejía R; Juaristi E Calorimetric and Computational Study of 1,4-Dithiacyclohexane 1,1-Dioxide (1,4-Dithiane Sulfone). J. Org. Chem 2006, 71, 2581–2586. [DOI] [PubMed] [Google Scholar]
- (88).Roux MV; Temprado M; Jiménez P; Notario R; Guzmán-Mejía R; Juaristi E Calorimetric and Computational Study of 1,3-Dithiacyclohexane 1,1-Dioxide (1,3-Dithiane Sulfone). J. Org. Chem 2004, 69, 1670–1675. [DOI] [PubMed] [Google Scholar]
- (89).Roux MV; Temprado M; Jiménez P; Notario R; Guzmán-Mejía R; Juaristi E Calorimetric and Computational Study of 1,3- and 1,4-Oxatiane Sulfones. J. Org. Chem 2007, 72, 1143–1147. [DOI] [PubMed] [Google Scholar]
- (90).Freitas VLS; Gomes JRB; Ribeiro da Silva MDMC Energetics and Reactivity of Morpholine and Thiomorpholine: A Joint Experimental and Computational Study. J. Chem. Eng. Data 2014, 59, 312–322. [Google Scholar]
- (91).Silva ALR; Monte MJS; Morais VMF; Ribeiro da Silva MDMC Thermodynamic Study of 2-Aminothiazole and 2-Aminobenzothiazole: Experimental and Computational Approaches. J. Chem. Thermodyn 2014, 74, 67–77. [Google Scholar]
- (92).Oliveira TSM; Freitas VLS; Ribeiro da Silva MDMC Energetic Insights on Two Dye Key Molecules: N-methylphenothiazine and N-methylphenoxazine. J. Chem. Thermodyn 2016, 94, 7–15. [Google Scholar]
- (93).Szterner P; Galvão TLP; Amaral LMPF; Ribeiro da Silva MDMC; Ribeiro da Silva MAV 5-Isopropylbarbituric and 2-Thiobarbituric Acids: An Experimental and Computational Study. Thermochim. Acta 2016. , 625, 36–46. [Google Scholar]
- (94).Silva ALR; Goncalves JM; Ribeiro da Silva MDMC Experimental and Computational Thermochemical Study of Two Fluorobenzazoles: 5-Fluoro-2-methylbenzoxazole and 5-Fluoro-2-methylbenzothiazole. J. Chem. Thermodyn 2018, 120, 157–163. [Google Scholar]
- (95).Camarillo EA; Mentado J; Flores H; Hernández-Pérez JM Standard Enthalpies of Formation of 2-Aminobenzothiazoles in the Crystalline Phase by Rotating-Bomb Combustion Calorimetry. J. Chem. Thermodyn. 2014, 73, 269–273. [Google Scholar]
- (96).Ramos F; Flores H; Rojas A; Hernández-Pérez JM; Camarillo EA; Amador MP Experimental and Computational Thermochemical Study of Benzofuran, Benzothiophene and Indole Derivatives. J. Chem. Thermodyn 2016, 97, 297–306. [Google Scholar]
- (97).Mackle H; O’Hare PAG Studies in the Thermochemistry of Sulphones. Part 5. – Heats of Combustion, Formation, Fusion, Vaporization and Atomization of Eleven Aliphatic Sulphones. Trans. Faraday Soc 1961, 57, 1070–1074. [Google Scholar]
- (98).Good W; Lacina J; McCullough J Methanethiol and carbon disulfide: heats of combustion and formation by rotating-bomb Calorimetry. J. Phys. Chem 1961, 65, 2229–2231. [Google Scholar]
- (99).McCullough JP; Hubbard WN; Frow FR; Hossenlopp IA; Waddington G Ethanethiol and 2-Thiapropane: Heats of Formation and Isomerization; The Chemical Thermodynamic Properties from 0 to 1000°K. J. Am. Chem. Soc 1957, 79, 561–566. [Google Scholar]
- (100).Hubbard WN; Waddington G The heats of combustion, formation and isomerization of propanethiol-1, propane-thiol-2 and 2-thiabutane. Recl. Trav. Chim. Pays-Bas 1954, 73, 910–923. [Google Scholar]
- (101).McCullough JP; Finke HL; Scott DW; Gross ME; Messerly JF; Pennington RE; Waddington G 2-Propanethiol: Experimental Thermodynamic Studies from 12 to 500°K. The Chemical Thermodynamic Properties from 0 to 1000°K. J. Am. Chem. Soc 1954, 76, 4796–4802. [Google Scholar]
- (102).Hubbard WN; Good WD; Waddington G The Heats of Combustion, Formation and Isomerization of the Seven Isomeric C4H10S Alkane Thiols and Sulfides. J. Phys. Chem 1958, 62, 614–617. [Google Scholar]
- (103).McCullough JP; Finke HL; Scott DW; Pennington RE; Gross ME; Messerly JF; Waddington G 2-Butanethiol: Chemical Thermodynamic Properties between 0 and 1000°K. Rotational Conformations. J. Am. Chem. Soc 1958, 80, 4786–4793. [Google Scholar]
- (104).McCullough JP; Scott DW; Finke HL; Hubbard WN; Gross ME; Katz C; Pennington RE; Messerly JF; Waddington G The Thermodynamic Properties of 2-Methyl-2propanethiol from 0 to 1000°K. J. Am. Chem. Soc 1953. , 75, 1818–1824. [Google Scholar]
- (105).Berg WT; Scott DW; Hubbard WN; Todd SS; Messerly JF; Hossenlopp IA; Osborn A; Douslin DR; McCullough JP The Chemical Thermodynamic Properties of Cyclopentanethiol. J. Phys. Chem 1961, 65, 1425–1430. [Google Scholar]
- (106).Scott DW; McCullough JP; Hubbard WN; Messerly JF; Hossenlopp IA; Frow FR; Waddington G Benzenethiol: Thermodynamic Properties in the Solid, Liquid and Vapor States; Internal Rotation of the Thiol Group. J. Am. Chem. Soc 1956, 78, 5463–5468. [Google Scholar]
- (107).Moravetz E; Sunner S Design, Construction, and Testing of a Heat of Vaporization Calorimeter Useful in the Vapor Pressure Range 1 to 0.01 mm Hg at 25°C. Acta Chem. Scand 1963, 17, 473–488. [Google Scholar]
- (108).Scott DW; Finke HL; McCullough JP; Messerly JF; Pennington RE; Hossenlopp IA; Waddington G 1-Butanethiol and 2-Thiapentane. Experimental Thermodynamic Studies between 12 and 500°K.; Thermodynamic Functions by a Refined Method of Increments. J. Am. Chem. Soc 1957, 79, 1062–1068. [Google Scholar]
- (109).Sapei E; Uusi-Kyyny P; Keskinen KI; Pokki J; Alopaeus V Phase Equilibria on Five Binary Systems Containing 1-Butanethiol and 3-Methylthiophene in Hydrocarbons. Fluid Phase Equilib. 2010, 293, 157–163. [Google Scholar]
- (110).Scott DW; McCullough JP; Messerly JF; Pennington RE; Hossenlopp IA; Finke HL; Waddington G 2-Methyl-1-propanethiol: Chemical Thermodynamic Properties and Rotational Isomerism. J. Am. Chem. Soc 1958, 80, 55–59. [Google Scholar]
- (111).Osborn AG; Douslin DR Vapor Pressure Relations of 36 Sulfur Compounds Present in Petroleum. J. Chem. Eng. Data 1966, 11, 502–509. [Google Scholar]
- (112).Hossenlopp IA; Scott DW Vapor Heat Capacities and Enthalpies of Vaporization of Six Organic Compounds. J. Chem. Thermodyn 1981, 13, 405–414. [Google Scholar]
- (113).Mackle H; McClean RTB Studies in the Thermochemistry of Organic Sulphides. Part 4.– Heat of Formation of the Mercaptyl Radical. Trans. Faraday Soc 1962, 58, 895–899. [Google Scholar]
- (114).Steele WV; Chirico RD; Hossenlopp IA; Knipmeyer SE; Nguyen A; Smith NK DIPPR Project 871. Determination of Ideal-Gas Enthalpies of Formation for Key Compounds. The 1990 Project Results. DIPPR Data Ser. 1994, 2, 188–215. [Google Scholar]
- (115).Månsson M; Sunner S Heats of Combustion and Formation of the n-Alkane-α, ω-dithiols from Ethane through Pentane. Acta Chem. Scand 1962, 16, 1863–1869. [Google Scholar]
- (116).Voronkov MG; Klyuchnikov VA; Kolabin SN; Shvets GN; Varushin PI; Deryagina EN; Korchevin NA; Tsvetnitskaya SI Thermochemical Properties of Diorganylchalkogenides and Dilchalkogenides RMnR (M = S, Se, Te; n = 1, 2). Dokl. Akad. Nauk SSSR 1989, 307, 1139–1143. [Google Scholar]
- (117).Hubbard WN; Katz C; Waddington G A Rotating Combustion Bomb for Precision Calorimetry. Heats of Combustion of Some Sulfur-Containing Compounds. J. Phys. Chem 1954, 58, 142–152. [Google Scholar]
- (118).Sunner S Corrected Heat of Combustion and Formation Values for a Number of Organic Sulphur Compounds. Acta Chem. Scand 1963, 17, 728–730. [Google Scholar]
- (119).Scott DW; Finke HL; Hubbard WN; McCullough JP; Katz C; Gross ME; Messerly JF; Pennington RE; Waddington G Thiacyclobutane: Heat Capacity, Heats of Transition, Fusion and Vaporization, Vapor Pressure, Entropy, Heat of Formation and Thermodynamic Functions. J. Am. Chem. Soc 1953, 75, 2795–2800. [Google Scholar]
- (120).Scott DW; Finke HL; McCullough JP; Gross ME; Williamson KD; Waddington G; Huffman HM Thermodynamic Properties and Rotational Isomerism of 2-Thiabutane1. J. Am. Chem. Soc 1951, 73, 261–265. [Google Scholar]
- (121).Wadsö I The Heats of Hydrolysis of Some Alkyl Thiolesters. Acta Chem. Scand 1957, 11, 1745–1751. [Google Scholar]
- (122).Wadsö I Heats of Vaporization for a Number of Organic Compounds at 25°C. Acta Chem. Scand 1966, 20, 544–552. [Google Scholar]
- (123).McCullough JP; Finke HL; Messerly JF; Pennington RE; Hossenlopp IA; Waddington G 3-Methyl-2-thiabutane: Calorimetric Studies from 12 to 500°K.; the Chemical Thermodynamic Properties from 0 to 1000°K. J. Am. Chem. Soc 1955, 77, 6119–6125. [Google Scholar]
- (124).McCullough JP; Finke HL; Hubbard WN; Good WD; Pennington RE; Messerly JF; Waddington G The Chemical Thermodynamic Properties of Thiacyclohexane from 0 to 1000°K. J. Am. Chem. Soc 1954, 76, 2661–2669. [Google Scholar]
- (125).Steele WV; Chirico RD; Cowell AB; Nguyen A; Knipmeyer SE, Possible Precursors and Products of Deep Hydrodesulfurization of Gasoline and Distillate Fuels. Part 2. The Thermodynamic Properties of 2,3-Dihydrobenzo[b]thiophene. J. Chem. Thermodyn 2003, 35, 1253–1276. [Google Scholar]
- (126).Mackle H; Mayrick RG, Studies in the Thermochemistry of Organic Sulphides. Part 2. – The Gas-Phase Heats of Formation of Ten Aliphatic Sulphides. Trans. Faraday Soc 1962, 58, 230–237. [Google Scholar]
- (127).Osborn AG; Scott DW, Vapor Pressures of 17 Miscellaneous Organic Compounds. J. Chem. Thermodyn 1980, 12, 429–438. [Google Scholar]
- (128).Mackle H; Mayrick RG, Studies in the Thermochemistry of Organic Sulphides. Part 1. – The Gas-Phase Heats of Formation of Phenyl Methyl, Phenyl Ethyl, Benzyl Methyl a nd Benzyl Ethyl Sulphides. Trans. Faraday Soc 1962, 58, 33–39. [Google Scholar]
- (129).Sabbah R; Xu-wu A; Chickos JS; Planas Leitão ML; Roux MV; Torres LA, Reference Materials for Calorimetry and Differential Thermal Analysis. Thermochim. Acta 1999, 331, 93–204. [Google Scholar]
- (130).Monte MJS; Sousa CAD; Fonseca JMS; Santos LMNBF, Thermodynamic Study on the Sublimation of Anthracene-Like Compounds. J. Chem. Eng. Data 2010, 55, 5264–5270. [Google Scholar]
- (131).Edwards DR; Prausnitz JM, Vapor Pressures of Some Sulfur-Containing, Coal-Related Compounds. J. Chem. Eng. Data 1981, 26, 121–124. [Google Scholar]
- (132).Sabbah R; El Watik L, Etude Thermodynamique de la Molecule de Thianthrene. Thermochim. Acta 1989, 138, 241–247. [Google Scholar]
- (133).Sandman DJ; Epstein AJ; Chickos JS; Ketchum J; Fu JS; Scheraga HA, Crystal Lattice and Polarization Energy of Tetrathiafulvalene. J. Chem. Phys 1979, 70, 305–313. [Google Scholar]
- (134).Steele WV; Chirico RD; Cowell AB; Nguyen A; Knipmeyer SE, Possible Precursors and Products of Deep Hydrodesulfurization of Distillate Fuels: I. The Thermodynamic Properties of Diphenylsulfide and Densities and Revised Properties for Dibenzothiophene. J. Chem. Thermodyn 1995, 27, 1407–1428. [Google Scholar]
- (135).Mackle H; Mayrick RG, Studies in the Thermochemistry of Organic Sulphides. Part 3. – The Gas-Phase Heats of Formation of Diphenyl Sulphide, Dibenzyl Sulphide, Diphenyl Disulphide, 4-Thia-1-hexene and 4-Thia-5,5-dimethyl-1-hexene. Trans. Faraday Soc 1962, 58, 238–243. [Google Scholar]
- (136).Sawaya T; Mokbel I; Rauzy E; Saab J; Berro C; Jose J, Experimental Vapor Pressures of Alkyl and Aryl Sulfides: Prediction by a Group Contribution Method. Fluid Phase Equilib. 2004, 226, 283–288. [Google Scholar]
- (137).Steele WV; Chirico RD; Knipmeyer SE; Nguyen A; Smith NK, Thermodynamic Properties and Ideal-Gas Enthalpies of Formation for Dicyclohexyl Sulfide, Diethylenetriamine, Di-n-octyl Sulfide, Dimethyl Car bonate, Piperazine, Hexachloroprop-1-ene, Tetrakis(dimethylamino)ethylene, N,N’-Bis-(2-hydroxyethyl)ethylenediamine, and 1,2,4-Triazolo[1,5-α]pyrimidine. J. Chem. Eng. Data 1997, 42, 1037–1052. [Google Scholar]
- (138).Steele WV; Chirico RD; Knipmeyer SE; Nguyen A, Possible Precursors and Products of Deep Hydrodesulfurization of Gasoline and Distillate Fuels IV. Heat Capacities, Enthalpy Increments, Vapor Pre ssures, and Derived Thermodynamic Functions for Dicyclohexylsulfide Between the Temperatures (5 and 520) K. J. Chem. Thermodyn 2004, 36, 845–855. [Google Scholar]
- (139).Davies JV; Sunner S, Heats of Combustion and Formation of Thiolane and the Thiolenes. Acta Chem. Scand 1962, 16, 1870–1876. [Google Scholar]
- (140).Hubbard WN; Finke HL; Scott DW; McCullough JP; Katz C; Gross ME; Messerly JF; Pennington RE; Waddington G Thiacyclopentane: Heat Capacity, Heats of Fusion and Vaporization, Vapor Pressure, Entropy, Heat of Formation and Thermodynamic Functions. J. Am. Chem. Soc 1952, 74, 6025–6030. [Google Scholar]
- (141).Scott DW; Good WD; Todd SS; Messerly JF; Berg WT; Hossenlopp IA; Lacina JL; Osborn A; McCullough JP, 3,3-Dimethyl-2-Thiabutane: Chemical Thermodynamic Properties and Barriers to Internal Rotation. J. Chem. Phys 1962, 36, 406–412. [Google Scholar]
- (142).Scott DW; Finke HL; McCullough JP; Gross ME; Pennington RE; Waddington G, 3,4-Dithiahexane: Heat Capacity, Heats of Fusion and Vaporization, Vapor Pressure, Entropy, and Thermodynamic Functions. J. Am. Chem. Soc 1952, 74, 2478–2483. [Google Scholar]
- (143).Kusano K, Micro Conduction Calorimeters to Measure Enthalpies of Vaporization. Thermochim. Acta 1985, 88, 109–120. [Google Scholar]
- (144).Mackle H; McClean RTB, Studies in the Thermochemistry of Organic Sulphides. Part 5. – Dithia-alkanes. Trans. Faraday Soc 1964, 60, 669–672. [Google Scholar]
- (145).Metzger RM; Kuo CS; Arafat ES, A Semi-Micro Rotating-Bomb Combustion Calorimeter. J. Chem. Thermodyn 1983, 15, 841–851. [Google Scholar]
- (146).Maånson M; Sunner S, Heats of Combustion and Formation of 4-Methylthiazole and 4-Cyanothiazole. Acta Chem. Scand 1966, 20, 845–848. [Google Scholar]
- (147).Hubbard WN; Scott DW; Frow FR; Waddington G, Thiophene: Heat of Combustion and Chemical Thermodynamic Properties. J. Am. Chem. Soc 1955, 77, 5855–5857. [Google Scholar]
- (148).Pennington RE; Finke HL; Hubbard WN; Messerly JF; Frow FR; Hossenlopp IA; Waddington G, The Chemical Thermodynamic Properties of 2-Methylthiophene. J. Am. Chem. Soc 1956, 78, 2055–2060. [Google Scholar]
- (149).Steele WV; Chirico RD; Knipmeyer SE; Nguyen A, The Thermodynamic Properties of Benzothiazole and Benzoxazole. J. Chem. Thermodyn 1992. , 24 , 499–529. [Google Scholar]
- (150).Sabbah R; Hevia R, Energétique des Liaisons Intermoléculaires dans les Molécules de Benzoxazole et de Benzothiazole. Thermochim. Acta 1998, 313, 131–136. [Google Scholar]
- (151).Ribeiro da Silva MAV; Santos AFLOM; Gomes JRB; Roux MV; Temprado M; Jiménez P; Notario R, Thermochemistry of Bithiophenes and Thienyl Radicals. A Calorimetric and Computational Study. J. Phys. Chem. A 2009, 113, 11042–11050. [DOI] [PubMed] [Google Scholar]
- (152).Chirico RD; Knipmeyer SE; Nguyen A; Steele WV, The Thermodynamic Properties of Benzo[b]thiophene. J. Chem. Thermodyn 1991, 23, 759–779. [Google Scholar]
- (153).Sabbah R; Antipine I; Coten M; Davy L, Quelques Reflexions a Propos de la Mesure Calorimetrique de l’Enthalpie de Sublimation ou Vaporisation. Thermochim. Acta 1987, 115, 153–165. [Google Scholar]
- (154).Sabbah R Thermodynamique de Substances Soufrees. I. Etude Thermochimique de Benzo-2,3-Thiophene et du Dibenzothiophene. Bull. Soc. Chim. Fr 1979, I–434–I–437. [Google Scholar]
- (155).Freitas VLS; Gomes JRB; Ribeiro da Silva MDMC, Revisiting Dibenzothiophene Thermochemical Data: Experimental and Computational Studies. J. Chem. Thermodyn 2009, 41, 1199–1205. [Google Scholar]
- (156).Sabbah R; Antipine I, Etude Thermodynamique Comparée de Quelques Substances Polycycliques. Recherche du Lien Existant Entre Leurs Grandeurs Énergétiques et Leur Structure. Bull. Soc. Chim. Fr 1987, 392–400. [Google Scholar]
- (157).McCullough JP; Sunner S; Finke HL; Hubbard WN; Gross ME; Pennington RE; Messerly JF; Good WD; Waddington G, The Chemical Thermodynamic Properties of 3-Methylthiophene from 0 to 1000°K. J. Am. Chem. Soc 1953, 75, 5075–5081. [Google Scholar]
- (158).Steele WV; Chirico RD; Cowell AB; Nguyen A; Knipmeyer SE, Possible Precursors and Products of Deep Hydrodesulfurization of Gasoline and Distillate Fuels III. The Thermodynamic Properties of 1,2,3,4-Tetrahydrodibenzothiophene. J. Chem. Thermodyn 2004, 36, 497–509. [Google Scholar]
- (159).Inagaki S; Murata S; Sakiyama M, Thermochemical Studies on Thioacetamide and Tetramethylthiourea. Estimation of Stabilization Energies Due to Interaction between Thiocarbonyl Group and Neighbouring Nitrogen Atom. Bull. Chem. Soc. Jpn 1982, 55, 2808–2813. [Google Scholar]
- (160).Sabbah R; Torres Gomez LA, Thermodynamique de Substances Soufrees. III. Etude Thermochimique de la Thioacetamide et de la Thiobenzamide. Thermochim. Acta 1982, 52, 285–295. [Google Scholar]
- (161).Mentado J; Flores H; Amador P, Combustion Energies and Formation Enthalpies of 2-SH-Benzazoles. J. Chem. Thermodyn 2008, 40, 1106–1109. [Google Scholar]
- (162).Fulem M; Ružička K; Ružička MK, Recommended Vapor Pressures for Thiophene, Sulfolane, and Dimethyl Sulfoxide. Fluid Phase Equilib. 1999, 303, 205–216. [Google Scholar]
- (163).Steele WV; Chirico RD; Knipmeyer SE; Nguyen A, Vapor Pressure, Heat Capacity, and Density Along the Saturation Line, Measurements for Cyclohexanol, 2-Cyclohexen-1-one, 1,2-Dichloropropane, 1,4-Di-tert-butylbenzene, (±)-2-Ethylhexanoic Acid, 2-(Methylamino)ethanol, Perfluoro-n-heptane, and Sulfolane. J. Chem. Eng. Data 1997, 42, 1021–1036. [Google Scholar]
- (164).Roux MV; Temprado M; Jiménez P; Foces-Foces C; Notario R; Parameswar AR; Demchenko AV; Chickos JS; Deakyne CA; Liebman JF, Experimental and Theoretical Study of the Structures and Enthalpies of Formation of 3H-1,3-Benzoxazole-2-thione, 3H-1,3-Benzothiazole-2-thione, and Their Tautomers. J. Phys. Chem. A 2010, 114, 6336–6341. [DOI] [PubMed] [Google Scholar]
- (165).Roux MV; Notario R; Zaitsau DH; Emel’yanenko VN; Verevkin SP, Experimental and Computational Thermochemical Study of 2-Thiobarbituric Acid: Structure-Energy Relationship. J. Phys. Chem. A 2012, 116, 4639–4645. [DOI] [PubMed] [Google Scholar]
- (166).Brunetti B; Piacente V, Sublimation Enthalpies Study for Barbituric, Tiobarbituric, and Selenobarbituric Acids from Their Vapor Pressure Measurements. J. Chem. Eng. Data 1999, 44, 809–812. [Google Scholar]
- (167).Butler KH; Maass O, Hydrogen Disufide. J. Am. Chem. Soc 1930, 52, 2184–2198. [Google Scholar]
- (168).Monte MJS; Santos LMNBF; Sousa CAD; Fulem M Vapor Pressures of Solid and Liquid Xanthene and Phenoxathiin from Effusion and Static Studies. J. Chem. Eng. Data 2008, 53, 1922–1926. [Google Scholar]
- (169).Goldfarb JL; Suuberg EM, Vapor Pressures and Sublimation Enthalpies of Seven Heteroatomic Aromatic Hydrocarbons Measured Using the Knudsen Effusion Technique. J. Chem. Thermodyn 2010, 42, 781–786. [DOI] [PMC free article] [PubMed] [Google Scholar]
