Next Article in Journal
Protective Effect of Baicalin on Chlorpyrifos-Induced Liver Injury and Its Mechanism
Next Article in Special Issue
Step-by-Step Replacement of Cyano Groups by Tricyanovinyls—The Influence on the Acidity
Previous Article in Journal
Harnessing the Power of Microwave Irradiation: A Novel Approach to Bitumen Partial Upgrading
Previous Article in Special Issue
On Prototropy and Bond Length Alternation in Neutral and Ionized Pyrimidine Bases and Their Model Azines in Vacuo
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Theoretical Study on Fluorinated Derivatives of Sulfolane, Cyclopentanone, and Gamma-Butyrolactone

Institute of Chemistry, University of Tartu, Ravila 14a, 50411 Tartu, Estonia
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(23), 7770; https://doi.org/10.3390/molecules28237770
Submission received: 29 August 2023 / Revised: 19 November 2023 / Accepted: 23 November 2023 / Published: 25 November 2023
(This article belongs to the Special Issue Computational and Theoretical Studies on Isomeric Organic Compounds)

Abstract

:
In this paper, fluorinated compounds based on sulfolane, cyclopentanone, and gamma-butyrolactone are studied computationally, focusing on their applicability in electrochemical devices and acid–base-related studies. Candidates for solvents with (1) high polarity, (2) good electrochemical stability, and (3) low basicity were searched for. Some of the compounds are studied here for the first time. Electrochemical stabilities, dielectric constants, boiling points, basicities, and lipophilicities were estimated using DFT and COSMO-RS methods with empirical corrections. The effect of fluorination on these properties as well as the bond parameters was studied. The possible synthesis routes of the proposed compounds are outlined. Some molecules display a combination of estimated properties favorable for a solvent, although none of the studied compounds are expected to surpass acetonitrile and propylene carbonate by the width of the electrochemical stability window.

Graphical Abstract

1. Introduction

Despite there being many solvents already in use for electrochemical applications [1,2], the search for new solvents and solvent additives is ongoing [3,4,5]. There are a number of properties that a compound must possess to be practically usable in electrochemical devices (batteries or supercapacitors) as a main solvent [2,5,6]:
High redox stability (voltage durability), as the capacity of a double-layer capacitor is proportional to the square of the operating voltage.
Low viscosity, which is a prerequisite for high ionic conductivity [7].
High ion solubilities to produce electrolytes with sufficient ion concentrations (achievable by the high free energy of the solvation of relevant ions and high dielectric constant).
Low enough melting point and high boiling and flash points.
Sufficient thermal stability.
Low protonation/deprotonation ability and low reactivity.
Additionally, low toxicity, low cost, and recyclability are desirable, if not strictly necessary. Compatibility with the electrode materials also plays a role, as some compounds that perform well with metal anodes may be incompatible with graphite anodes or vice versa.
Most properties that make an excellent electrolyte solvent are also needed in studies of strong acids and creating strongly acidic systems: high dielectric constant and solvation ability prevent undesirable interactions between ions, and a low reactivity allows for a wider choice of components. Low basicity is a necessary property for the study of strong acids, as the basic nature of the solvent has a leveling effect on the dissociation of acids and hinders differentiating their strengths. A sufficiently high boiling point is important for safety and retaining stable concentrations.
An ideal solvent hardly exists, as some of the listed properties are correlated in such a way that achieving two within the optimal range is difficult. The solvent’s ability for specific solvation, on the one hand, promotes salt solubility and ion separation; on the other hand, it usually comes with more pronounced acid–base properties and higher viscosity. For these reasons, protic solvents are generally unsuitable for the above-mentioned purposes. Most solvents currently used in nonaqueous electrochemistry are dipolar aprotic: carbonates, ethers, esters, and nitriles [1,2,8].
Fluorination can enhance some of the desirable characteristics of solvents, which, however, comes at the cost of some simultaneous undesirable changes. A variety of fluorinated compounds have been explored as solvents or electrolyte components for electrochemical devices, either experimentally [8,9,10,11,12,13,14] or theoretically [15,16,17]. Compared to nonfluorinated analogs, partially fluorinated compounds tend to have the following:
Higher oxidation resistance due to a lower HOMO energy [9,18], but also a higher tendency to be reduced due to a lower LUMO energy [9,15].
Lower flammability or even fire-suppressing properties [8,19,20].
Higher dielectric constant, if the F atom is placed in the right position to increase the dipole moment of the molecule.
Lower basicity and higher acidity.
Higher viscosity due to an increase in molecular mass and hydrogen bond donicity [10,11,13], although, in some cases, viscosity reduction has been reported as well [21].
Most of the time, fluorinated (especially polyfluorinated) compounds are less reactive than their nonfluorinated analogs. However, in some cases, fluorination increases reactivity [15,16,22]. For example, alpha-fluorination weakens C-S bonds in sulfones [15,16], which contributes to the increase in reduction potential.
In this work, we choose three scaffold compounds—sulfolane (SL), cyclopentanone (CP), and gamma-butyrolactone (GBL)—to be modified by fluorination and the addition of double bonds and characterize the obtained molecules in the context of electrochemistry and studies of strong acids. The studied structures are shown in Figure 1.
Of the parent compounds, SL and GBL are used as electrolyte components [1,8]. SL has been suggested as facilitating the fast transport of lithium ions through the unique hopping diffusion mechanism [23].
The available literature data (Table 1) show that adding a double bond to a nonfluorinated parent structure dramatically increases the melting point (MP). Since the MP depends strongly on the compound purity, we abstained from making quantitative conclusions. However, it is clear that many compounds containing a double bond may not be usable as pure solvents but only in combination with lower-melting compounds. The effect of fluorination on the MP is not obvious since reduced symmetry and increased polarity have opposite influences on the MP. There are examples of both an increase and decrease in MP upon the addition of F atom(s) [13]. Considering the very low MP of the parent CP and GBL, it is likely that the solvents of the series CP-A and GBL-A are liquid at room temperature and usable in pure form.
Another potentially problematic aspect is the reactivity of the unsaturated compounds, which can limit their usability in harsh conditions and in combination with other reactive species. Most solvents that are typically used in electrochemical devices are composed of saturated molecules because double bonds compromise stability. However, unsaturated compounds can be used as electrolyte additives. They may be required to have a lower stability than the main solvent in order to react and form a film on the electrode surface [32,33,34].
The properties explored in this study are:
Electrochemical stability windows (ESWs) were estimated via the calculated oxidation and reduction potentials in gas phase and DMSO, or approximated via the differences of the gas-phase HOMO and LUMO energies [35].
Dielectric constants (εr) and boiling points (BP) were computed with the COSMO-RS method with empirical corrections.
Basicities were characterized via the predicted Gibbs energies of the transfer of a proton from water to the studied solvent (∆trG°(H+)), basicity in acetonitrile (pKaH(MeCN), corresponding to the pKa of the protonated molecule), and gas-phase basicity (GB) of the solvents.
Mutual solubility with water and lipophilicity were computed with the COSMO-RS method.
εr values, ∆trG°(H+) values, and solubilities were computed on the assumption that the compounds are liquid at room temperature (which, as explained above, this may not always be the case).

2. Results

The results of the computations, together with some values from the literature, are presented in Table 2.

3. Discussion

3.1. Effect of Fluorination on Bond Strengths

The bond parameters were analyzed in a similar manner to ref. [16]. As expected, bond lengths were found to be well correlated with the bond strengths within compound groups. The estimated changes in bond strengths (expressed as intrinsic bond strength indices (IBSI) [41]) and multiplicities (expressed as Wiberg bond indices (WBI) [42]) with fluorination are shown in Figure 2, Figure 3 and Figure 4. All bond length, WBI, and IBSI values are provided in Table S5.
It was found that, in all studied compound groups, α-fluorination weakens the bond between the fluorinated carbon and electronegative group and reduces its bond order. At the same time, it can strengthen other bonds. Also, a slight strengthening of the C=O and S=O bonds was observed. In sulfolanes (Figure 2), the bond weakening upon fluorination is the most pronounced, and the effect of the second F atom is similar to the effect of the first. This is in agreement with the literature [16]. The strengthening of the C2-C3 bond, especially noticeable when the bond is single, is indicated by the increased IBSI; however, this is at odds with the reduced WBI of the bond.
Similar effects were observed for cyclopentanone derivatives (Figure 3). α-fluorination weakens the C1-C2 bond, while slightly strengthening the C1-C5 bond and possibly the C2-C3 bond. Fluorination at C5 weakens the C1-C5 bond, while slightly strengthening the C1-C2 and C4-C5 bonds. The effect of the second F atom on bond weakening is mostly similar to that of the first F atom, but its effect on bond strengthening tends to be weaker.
In GBL derivatives (Figure 4), α-fluorination weakens the bond between C2 and C1 (carbonyl), while fluorination at C4 strengthens the C4-O5 bond but weakens the C1-O5 bond. The increase in the IBSI of C4-O5 is outstanding compared to the SL and CP derivatives, being over 10% for subgroup A (no double bonds). The effect of the second F atom is mostly weaker than that of the first.

3.2. Effect of Structural Variations on the Electrochemical Stability

The calculated oxidation and reduction free energies in the gas phase and DMSO (as an example of a polar aprotic environment) are presented in Figure 5. The values are presented relative to those of sulfolane (SL-A0). For several compounds, the geometries of the oxidized/reduced species were broken up or rearranged during the geometry optimization step of the computation, hinting at the low stability of these species (such geometries were not used for further computations). It can be seen that, first, solvation in DMSO does not significantly change the stability trends compared to the gas phase (Figure 5), and second, HOMO and LUMO levels obtained with the B3LYP functional adequately describe the trends in stability (Figure 6).
The computations show that fluorination (both number and position of F atoms) has a relatively small effect on the compound’s propensity to oxidation or reduction, compared to the significant drops in reduction stability that accompany the additions of the double bond(s) adjacent to the polar group (series B and D, and also C for GBL derivatives). The structural variations in the studied molecules influence reduction more than oxidation. Most studied molecules have lower reduction stabilities than MeCN, PC, and DMSO. None of the compounds have resistance to oxidation that is superior to MeCN, although in this respect, many outperform DMSO, and several GBL group compounds outperform PC.
As the addition of fluorine tends to lower both HOMO and LUMO levels in the studied compounds, the net effect of fluorination can be both positive and negative, depending on the structure. In most studied cases, however, fluorination slightly narrowed ESW.

3.3. Effect of Structural Variations on Other Properties

As seen in Figure 7 and Table 2, the effect of fluorination on compound properties varies according to compound group. The dielectric constant increases in the CP group but decreases in the GBL group. The likely reason is the different effects of fluorination on the dipole moment of the molecules. In the studied CP derivatives, fluorination always increases the dipole moment, while in GBL, the dipole moment can change both ways depending on substitution. As in the case of bond strengths, the effect of F atoms is not additive. The changes in εr values in the SL group are within the prediction uncertainty. A similar situation is observed with the boiling points of fluorinated compounds: an increase in CP, a decrease in GBL, and small and inconsistent changes in SL. ∆trG°(H+) values are invariably increased by fluorination, and the effect of F atoms appears to be more or less additive. Lipophilicities (expressed as octanol/water partition coefficients) are notably higher for fluorinated GBL and SL derivatives, but the effect is weaker and less consistent in the CP group. A more pronounced increase in the dipole moment of CP likely counters the lipophilicity-inducing effect of the F atoms. The patterns of pKaH(MeCN) match the inverted patterns of ∆trG°(H+). Each F atom decreases the basicity by 2.5–3 orders of magnitude.
The differences between the positional isomers of the fluorinated compounds were rather insignificant in SL and CP groups and only became notable in GBL-type scaffolds with double bonds (GBL-B1a vs. GBL-B1b, and GBL-C1a vs. GBL-C1b). While the differences in εr, BP, and logPo/w matched the differences in molecular dipoles (GBL-B1b and GBL-C1b have lower dipole moments), the differences in basicity can be attributed to differences in electron delocalization: GBL-B1b and GBL-C1a, where F is located at the sp3 carbon, are less basic than their isomers, where F is conjugated with O atoms.

3.4. Solubility Trends

Figure 8 demonstrates the predicted solubility values of the studied compounds in water and of water in the studied compounds. It must be stressed that, if the compound is solid at room temperature, then the absolute values may be significantly overestimated, as the computation does not take into account the Gibbs energy of fusion. However, these predictions successfully demonstrate the hydrophobic effect of the added fluorine atoms. The solubilities of the studied compounds in water are more affected by fluorination than the solubility of water in them. While fluorination may increase polarity, which in itself facilitates solubility in polar media, this effect is clearly weaker than that of the larger size and low polarizability of fluorine atoms.

3.5. Synthesis of the Proposed Compounds

Sulfolane (SL-A0), 1,1-dioxide-2,3-dihydro-thiophene (SL-B0), 3-sulfolene (SL-C0), and 1,1-dioxide-thiophene (SL-D0) are widely available compounds whose synthesis originates either from alkenes (with the insertion of SO2) [43], from thiophenes (generally using H2O2 as an oxidizer) [44,45,46], or by catalytic double bond migration procedures (with transition metal catalysts) [47,48]. Their fluorinated derivatives can therefore be synthesized in a similar way using fluorinated alkenes (SL-A2, SL-C2), thiophenes (SL-D1) [49], or fluorinated 1,1-dioxide-2,3-dihydro-thiophenes (SL-B1, SL-B2) as precursors. Additionally, fluorination using a gaseous F2/N2 mixture of parent compounds has been employed (SL-A1) [50]. Perhaps SL-C1 could be prepared similarly. The preparation of the tri-fluorinated analogs SL-A3, SL-B3, and SL-C3 could be challenging by the abovementioned methods because the respective fluorinated alkene and thiophene are not available and their preparation methods have not been published.
γ-Butyrolactone (GBL-A0) can be fluorinated using F2/N2 gas [51] to obtain a mixture of GBL-A1a and GBL-A1b. The deoxyfluorination of hydroxy-γ-butyrolactone is also a well-established procedure [52]. The preparation of GBL-A2 was not described, but the preparation could be considered from 2,2-difluorobutanedioic acid [53,54], 2,2-difluoro-1,4-butanediol by cyclization [55], or from 3,3-difluorodihydro-2,5-furandione [55] by hydrogenation. These precursors are readily available compounds. The fluorination of parent compounds similar to ethylene carbonate may be possible [56]. GBL-B1a [57] and GBL-B1b [58] are readily available compounds that can be prepared via cyclization. The preparation of GBL-B2 by the isomerization of fumaroyl fluoride seems to be rather complicated [59]. Methods for fluorinated 2(3H)-furanone (GBL-C0) derivatives have not been published; however, synthesis from fluorinated precursors or double bond migration could be employed [60].
Cyclopentanone (CP-A0) is fluorinated to obtain a mono-fluorinated derivative (CP-A1) using different fluorinating agents (Accufluor, XeF2, etc.) [61,62,63]. CP-A2 is a readily available (although expensive) compound. Its synthesis methods have not been published, but synthesis from difluoro adipic acid esters may be possible [64]. Cyclopentenone CP-B1a is prepared using F2/N2 gas [65]. Methods for other cyclopentenone derivatives, as well as methods for fluorinated cyclopentadienone (CP-D1), are absent in the literature. Using fluorinated precursors or different fluorination methods should be applicable to obtain some degree of fluorination.

4. Conclusions

The problem when choosing suitable solvents for nonaqueous applications involving electrochemistry and/or strong acidity is that many beneficial solvent properties are correlated with unfavorable ones. Therefore, trade-offs have to be found. Figure 9 illustrates the relative advantages of the studied compound groups in terms of the parameters that are essential for the main solvent: BP (well correlated with flash points [66]), ∆trG°(H+) values as estimates of solution basicity (higher values indicate lower basicities), and dielectric constant and HOMO-LUMO gap as a first estimate of electrochemical stability. As expected, sulfolanes (especially subgroups A and B) surpass the other groups in terms of a favorable combination of properties: ESWs are estimated to be wide, BP and εr are relatively high, and basicities are medium to low. CP are the least suitable main solvent candidates given their relatively high basicities, low boiling points, and εr and HOMO-LUMO gaps. Some of GBL derivatives (series A and B) look promising with wide ESW, high εr, and medium basicity.
The presence of double bonds (subgroups B–D) is expected to significantly increase melting points and decrease reduction stability, making the respective compounds less likely to be suited for the role of main solvent and more eligible for the role of reactive electrolyte additive.

5. Materials and Methods

5.1. Computational Parameters and Software

Gas-phase properties were computed with the Gaussian 16 Rev. A.03 software [67]. The geometries of all the conformers of uncharged species were created and optimized using DFT at the M06-2X/6-311+G** and B3LYP/6-311+G** levels of theory. The geometries of reduced/oxidized species were optimized at the M06-2X/6-311+G** level of theory. For the calculation of solvation free energies in DMSO, geometry optimization for uncharged and reduced/oxidized species was carried out at the M06-2X/6-31+G* level in the gas phase and using the SMD model [68] (solvent DMSO). For all optimized geometries, vibrational spectra were computed to ensure that the optimized geometries correspond to the true energy minima. Small imaginary frequencies in two calculations (compounds SL-C2 and CP-B2) could not be removed by reoptimization and were ignored.
Oxidation and reduction free energies (∆oxG and ∆redG) for molecule S in the gas phase were computed from the lowest-energy conformers at the M06-2X/6-311+G** level of theory, as follows:
oxG(S0) = G(S+)gasG(S0)gas
redG(S0) = G(S)gasG(S0)gas
Solvation free energies (∆solvG) were calculated as the differences of the free energies of the molecules in DMSO and in the gas phase, computed at M06-2X/6-31+G* (with and without the SMD model, respectively).
Oxidation and reduction free energies in DMSO were calculated as follows:
oxG(S0)DMSO = G(S+)gas + ∆solvG(S+) − G(S0)gas − ∆solvG(S0)
redG(S0)DMSO = G(S)gas + ∆solvG(S) − G(S0)gas − ∆solvG(S0)
The EHOMO and ELUMO values were obtained from the lowest energy conformers at the B3LYP/6-311+G** level of theory.
Wiberg bond indices (WBI) [42] and intrinsic bond strength indices (IBSI) [41] were computed with the Multiwfn software (version 3.8, in development, accessed on 9 June 2023) [69] using the results calculated by the B3LYP/6-311+G** method.
GB values were computed with the G4MP2 method [70].
The COSMO-RS method was used for computing εr, pKaH, ΔtrG°(H+), BP, logPo/w values, and solubilities. The geometries of all conformers and protomers of the studied solvents were optimized at the BP86/TZVP level of theory in an ideal conductor (COSMO model), followed by single-point energy calculation at the BP86/def2-TZVPD level of theory with the Fine cavity parameter. Vibrational spectra were computed to ensure that the obtained geometries correspond to energy minima. Solvent properties were computed from the obtained surface charge distributions using the COSMOtherm software (release 2023) [71] using all conformers of the involved species. DFT computations were carried out using Turbomole V6.5 software [72].

5.2. Solvation Models

COSMO-RS (Conductor-Like Solvation Model for Real Solutions) [73,74,75,76] is a method based on DFT computations and statistical thermodynamics that can be used for modeling arbitrary multicomponent fluids. As a first step, the geometries of all involved species are optimized in the ideal conductor, which yields the partial charge distribution on the molecular surface and energy of the structure. In the second step, the interactions of the molecules in the fluid mixture are accounted for via the pair-wise interactions of their surface segments. This allows access to a variety of properties determined by the free energies of molecules in the studied media (liquid–vapor diagrams, partition coefficients, solvation energies, etc.). Unlike most other solvation models, COSMO-RS is parametrized at the atomic level and is suitable for modeling novel substances for which no experimental data exist to date.
SMD is an implicit solvation model based on solute electron density [68]. The solvent is represented as a dielectric continuum at the solute–solvent boundary. SMD provides solvation energies comparable in quality to those obtained with COSMO-RS, but is not usable for arbitrary liquids as its parametrization relies on the experimental data of the solvent.

5.3. Accuracy of the Computed Values and Applied Corrections

trG°(H+): This property is used as the basicity estimate of the solvents and is defined as the Gibbs energy change i the following process (Equation (5)):
(H3O+)H2O + (S)S ⇄ (H2O)H2O + (SH+)S
where S denotes the studied solvent and the subscripts denote the corresponding environment. The ∆trG°(H+) values of the process for several solvents were computed with the COSMO-RS method and correlated with the values from ref. [38] (details in Table S1). The results are shown in Figure 10. While the absolute errors of the computed values are significant, the experimental and computed values correlate satisfactorily within the compound groups with the same protonation center. The correction equation for the solvents of interest (Equation (6)) was composed using the six aprotic solvents protonating on the O atom. Standard deviations of the regression parameters are provided in parentheses. Sres refers to the overall regression standard deviation (standard deviation of the residuals). Given the large value of Sres, the predicted ∆trG°(H+) values can be considered only semiquantitative. The liquid state of the solvent was assumed in all calculations.
trG°Exp(H+) = ∆trG°Calc(H+) · 0.76(0.13) + 23.1(7.2)
N = 6, R2 = 0.89, Sres = 18 kJ mol−1
BP: The accuracy of the boiling point values computed with COSMO-RS was assessed using a set of solvents containing structural features similar to the studied compounds and covering the expected span of BP values (Figure 11, Table S2). Equation (7) was used to correct the values systematically overestimated by computations. The standard errors of the regression parameters are provided in parentheses. The estimated standard uncertainty of the corrected values was 18 K. The data show that the effect of fluorination is predicted acceptably well, considering the overall accuracy. The magnitude of the effect of adding double bonds, however, tends to be overestimated.
BPExp = BPCalc · 0.86(0.04) + 48(21)
N = 29, R2 = 0.94, Sres = 18 K
Free Energies of Oxidation and Reduction: The functional M06-2X has one of the best performances to predict the gas-phase ionization potentials and electron affinities amongst the popular DFT functionals, with mean unsigned errors as low as 10 and 13 kJ mol−1, respectively [77]. For the SMD model, the mean unsigned errors of the solvation free energies in DMSO computed at the M05-2X/6-31G* level of theory were estimated as 0.64 kcal mol−1 for neutrals and 4.5 kcal mol−1 for anions [68]. With the level of theory used in this work, we would expect the errors to be even lower. SMD was shown to perform decently in the prediction of aqueous reduction potentials, outperforming COSMO-RS for aprotic solutes [77].
Using HOMO and LUMO levels in this work relied on the following approximations [35,78]:
oxGIP ≈ −EHOMO
redGEA ≈ −ELUMO
where ∆oxG and ∆redG—free energy of oxidation/reduction, IP—ionization potential, and EA—electron affinity. EHOMO and ELUMO depend considerably on the computational method but correlate well with the oxidation and reduction potentials, respectively [3].
Dielectric constant (εr): The accuracy of dielectric constant estimates by the COSMO-RS method was assessed using the set of 20 solvents presented in Table S3. The solvents were selected by similarity to the compounds of interest: each contains one ring, no N atoms, and preferably one or more F atoms. The correction Equation (10) was obtained (standard deviation of regression parameters are shown in parentheses):
log10(εr, exp.) = log10(εr, calc.) · 1.12(0.05) + 0.01(0.05)
N = 20, R2 = 0.97, Sres = 0.08 log units
To verify the reliability of the approach, εr values of some structurally assorted solvents were predicted (Table S3, Figure S1). The results demonstrate that the predictions are mostly accurate at moderate εr ranges, but εr values of around or over 40 can be considerable in error and more likely under- than overestimated. The εr values for fluoropyridines and dimethylsulfate were strongly underestimated. For solvents structurally similar to the model compounds, we considered predictions as quantitative at εr < 35 (RMSE ca. 4) and semiquantitative at εr > 35 (RMSE > 10). The semiquantitative nature of the prediction at high εr values is not a problem because εr values above 35 are certainly suitable for the purposes of this study; thus, knowing the accurate εr value is not critically important.
pKaH: The pKa and pKaH values predicted by COSMO-RS are usually biased, yet well correlated with the respective experimental data. With a lack of suitable experimental data to evaluate or correct the basicity values in acetonitrile (pKaH(MeCN)), we suggest that these values be used for semiquantitative comparison between structurally similar compounds rather than to access absolute basicities.
logPo/w: COSMO-RS is known to predict the octanol–water partition coefficients very well, with some rare exceptions [79,80]. The standard error of the predictions is conservatively estimated as ca. 0.3 units; therefore, no corrections are needed. The values correspond to wet octanol at 25 °C.
Mutual Solubility with Water: The reported accuracy estimates for the aqueous solubility predictions by COSMO-RS vary by solute type and the parametrization used. The expected error for liquid solutes is 0.3 [81] log units or lower (0.1 log units in ref. [82]). However, for solid solutes, it is 0.5 [81] log units or higher [82]. Dupeux et al. [80] reported an RMSE of 0.18 log units for a set containing liquid and solid solutes; the higher accuracy may be partly due to the improved method parametrization. COSMO-RS predicted the solubility trends of water in various hydrocarbons quite well; the relative bias of the predicted mole fractions was within 29% [83]. For many studied compounds, we do not know with certainty whether they are liquid or solid at room temperature. The solubility estimates assume a liquid state of the solvents. The results were obtained using the “Liquid Extraction” option of COSMOtherm at 25 °C. No corrections were used.
Note on Viscosity Computations: COSMO-RS predicted the viscosities of a set of typical electrolyte solvents with a mean absolute deviation of 0.22 cP and a good correlation between the predictions and experimental data [35]. However, a test using the available data on fluorinated solvents and their nonfluorinated analogs showed that the effect of fluorination is not well reproduced by computations. It is possible that the random errors are of similar magnitude to the real effect of fluorination. Thus, viscosities were not reported in this work.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28237770/s1. Table S1: ∆trG°(H+) values presented in Figure 10; Table S2: Dataset used for assessment and correction of the calculated boiling points; Table S3: Dataset used for the correction of the calculated ε values; Figure S1: The experimental and predicted εr values; Table S4: The estimated properties of the studied molecules; Table S5: Bond parameters of the studied solvents. Publication [84] is cited in the SI.

Author Contributions

Conceptualization, I.L.; Investigation, S.T. and A.K.; Formal analysis, S.T., Visualization, S.T.; Writing—original draft, S.T. and A.K.; Writing—review and editing, A.K. and I.L., Funding acquisition, I.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Estonian Research Council, grant PRG690.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in Zenodo at dx.doi.org/10.5281/zenodo.10204051 (accessed on 22 November 2023).

Acknowledgments

The QC calculations were carried out in the High-Performance Computing Center of the University of Tartu [85].

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhong, C.; Deng, Y.; Hu, W.; Qiao, J.; Zhang, L.; Zhang, J. A Review of Electrolyte Materials and Compositions for Electrochemical Supercapacitors. Chem. Soc. Rev. 2015, 44, 7484–7539. [Google Scholar] [CrossRef] [PubMed]
  2. Li, M.; Hicks, R.P.; Chen, Z.; Luo, C.; Guo, J.; Wang, C.; Xu, Y. Electrolytes in Organic Batteries. Chem. Rev. 2023, 123, 1712–1773. [Google Scholar] [CrossRef] [PubMed]
  3. Cheng, L.; Assary, R.S.; Qu, X.; Jain, A.; Ong, S.P.; Rajput, N.N.; Persson, K.; Curtiss, L.A. Accelerating Electrolyte Discovery for Energy Storage with High-Throughput Screening. J. Phys. Chem. Lett. 2015, 6, 283–291. [Google Scholar] [CrossRef] [PubMed]
  4. Marcou, G.; Flamme, B.; Beck, G.; Chagnes, A.; Mokshyna, O.; Horvath, D.; Varnek, A. In Silico Design, Virtual Screening and Synthesis of Novel Electrolytic Solvents. Mol. Inform. 2019, 38, e1900014. [Google Scholar] [CrossRef] [PubMed]
  5. Schütter, C.; Husch, T.; Viswanathan, V.; Passerini, S.; Balducci, A.; Korth, M. Rational Design of New Electrolyte Materials for Electrochemical Double Layer Capacitors. J. Power Sources 2016, 326, 541–548. [Google Scholar] [CrossRef]
  6. Balducci, A. Electrolytes for High Voltage Electrochemical Double Layer Capacitors: A Perspective Article. J. Power Sources 2016, 326, 534–540. [Google Scholar] [CrossRef]
  7. Gong, K.; Fang, Q.; Gu, S.; Li, S.F.Y.; Yan, Y. Nonaqueous Redox-Flow Batteries: Organic Solvents, Supporting Electrolytes, and Redox Pairs. Energy Environ. Sci. 2015, 8, 3515–3530. [Google Scholar] [CrossRef]
  8. Zhang, J.; Zhang, J.; Liu, T.; Wu, H.; Tian, S.; Zhou, L.; Zhang, B.; Cui, G. Toward Low-Temperature Lithium Batteries: Advances and Prospects of Unconventional Electrolytes. Adv. Energy Sustain. Res. 2021, 2, 2100039. [Google Scholar] [CrossRef]
  9. Takehaea, M.; Ebara, R.; Nanbu, N.; Ue, M.; Sasaki, Y. Electrochemical Properties of Fluoro-γ-Butyrolactone and Its Application to Lithium Rechargeable Cells. Electrochemistry 2003, 71, 1172–1176. [Google Scholar] [CrossRef]
  10. Nambu, N.; Matsushita, Y.; Takehara, M.; Sasaki, Y. Physical and Electrochemical Properties of Fluorinated Dialkyl Ethers. Electrochemistry 2016, 84, 776–778. [Google Scholar] [CrossRef]
  11. Yue, Z.; Dunya, H.; Aryal, S.; Segre, C.U.; Mandal, B. Synthesis and Electrochemical Properties of Partially Fluorinated Ether Solvents for Lithiumsulfur Battery Electrolytes. J. Power Sources 2018, 401, 271–277. [Google Scholar] [CrossRef]
  12. Lu, H.; He, L.; Yuan, Y.; Zhu, Y.; Zheng, B.; Zheng, X.; Liu, C.; Du, H. Synergistic Effect of Fluorinated Solvents for Improving High Voltage Performance of LiNi0.5Mn1.5O4 Cathode. J. Electrochem. Soc. 2020, 167, 120534. [Google Scholar] [CrossRef]
  13. Todorov, Y.M.; Aoki, M.; Mimura, H.; Fujii, K.; Yoshimoto, N.; Morita, M. Thermal and Electrochemical Properties of Nonflammable Electrolyte Solutions Containing Fluorinated Alkylphosphates for Lithium-Ion Batteries. J. Power Sources 2016, 332, 322–329. [Google Scholar] [CrossRef]
  14. Zhang, S.S.; Read, J. Partially Fluorinated Solvent as a Co-Solvent for the Non-Aqueous Electrolyte of Li/Air Battery. J. Power Sources 2011, 196, 2867–2870. [Google Scholar] [CrossRef]
  15. Du, F.; Ye, T.; Shi, Y.; Zhang, Y.; Qiu, Z.; Liu, W.; Lv, T.; Duan, S.; Liu, J. Deciphering Reduction Stability of Sulfone and Fluorinated Sulfone Electrolytes: Insight from Quantum Chemical Calculations. Chem. Phys. 2023, 568, 111840. [Google Scholar] [CrossRef]
  16. Khamitov, E.M.; Kuz’mina, E.V.; Karaseva, E.V.; Kolosnitsyn, V.S. Evaluation of Electrochemical Stability of Substituted Sulfolanes Based on Bond Orders. Russ. J. Phys. Chem. A 2021, 95, 730–735. [Google Scholar] [CrossRef]
  17. Yang, X.; Liu, X.; Han, J.; Liu, Z.; Zhang, X. Tuning Fluorination of Carbonates for Lithium-Ion Batteries: A Theoretical Study. J. Phys. Chem. B 2023, 127, 3026–3040. [Google Scholar] [CrossRef] [PubMed]
  18. Khetan, A.; Pitsch, H.; Viswanathan, V. Identifying Descriptors for Solvent Stability in Nonaqueous Li–O2 Batteries. J. Phys. Chem. Lett. 2014, 5, 1318–1323. [Google Scholar] [CrossRef] [PubMed]
  19. Mathieu, O.; Diévart, P.; Turner, M.A.; Mohr, D.J.; Grégoire, C.M.; Alturaifi, S.A.; Catoire, L.; Petersen, E.L. Experimental and Detailed Kinetics Modeling Study of the Fire Suppressant Properties of di(2,2,2trifluoroethyl) Carbonate. Proc. Combust. Inst. 2023, 39, 499–510. [Google Scholar] [CrossRef]
  20. Yang, A.; Yang, C.; Xie, K.; Xin, S.; Xiong, Z.; Li, K.; Guo, Y.-G.; You, Y. Benchmarking the Safety Performance of Organic Electrolytes for Rechargeable Lithium Batteries: A Thermochemical Perspective. ACS Energy Lett. 2023, 8, 836–843. [Google Scholar] [CrossRef]
  21. Möller, K.-C.; Hodal, T.; Appel, W.K.; Winter, M.; Besenhard, J.O. Fluorinated Organic Solvents in Electrolytes for Lithium Ion Cells. J. Power Sources 2001, 97–98, 595–597. [Google Scholar] [CrossRef]
  22. Bryantsev, V.S.; Giordani, V.; Walker, W.; Blanco, M.; Zecevic, S.; Sasaki, K.; Uddin, J.; Addison, D.; Chase, G.V. Predicting Solvent Stability in Aprotic Electrolyte Li–Air Batteries: Nucleophilic Substitution by the Superoxide Anion Radical (O2•–). J. Phys. Chem. A 2011, 115, 12399–12409. [Google Scholar] [CrossRef]
  23. Ho, J.S.; Borodin, O.A.; Ding, M.S.; Ma, L.; Schroeder, M.A.; Pastel, G.R.; Xu, K. Understanding Lithium-ion Transport in Sulfolane- and Tetraglyme-Based Electrolytes Using Very Low-Frequency Impedance Spectroscopy. Energy Environ. Mater. 2023, 6, e12302. [Google Scholar] [CrossRef]
  24. Tilstam, U. Sulfolane: A Versatile Dipolar Aprotic Solvent. Org. Process Res. Dev. 2012, 16, 1273–1278. [Google Scholar] [CrossRef]
  25. Timmermans, J. The Melting Point of Organic Substances. X Bull. Soc. Chim. Belg. 1927, 36, 502–518. [Google Scholar]
  26. Ding, M.S.; Jow, T.R. Properties of PC-EA Solvent and Its Solution of LiBOB Comparison of Linear Esters to Linear Carbonates for Use in Lithium Batteries. J. Electrochem. Soc. 2005, 152, A1199. [Google Scholar] [CrossRef]
  27. Krug, R.C.; Tichelaar, G.R.; Didot, F.E. Unsaturated Cyclic Sulfones. III. Some Halogen-Containing Derivatives. J. Org. Chem. 1958, 23, 212–215. [Google Scholar] [CrossRef]
  28. Clauson-Kaas, N.; Elming, N.; Rottenberg, M.; Stenhagen, E.; Östling, S. 2-Acetoxyfuran. Acta Chem. Scand. 1952, 6, 560–564. [Google Scholar] [CrossRef]
  29. Kotha, S.; Khedkar, P.; Ghosh, A.K. Synthesis of Symmetrical Sulfones from Rongalite: Expansion to Cyclic Sulfones by Ring-Closing Metathesis. Eur. J. Org. Chem. 2005, 2005, 3581–3585. [Google Scholar] [CrossRef]
  30. Perrotti, E.; Palladino, N.; Greco, M.; De Malde, M. Reactions of Nitrosyl Salts. I. Reaction of Nitrosyl Sulfate with Succinic-Type Dicarboxylic Acids and 3,4-Dialkylisoxazol-5-Ones. Ann. Chim. 1966, 56, 1358–1372. [Google Scholar]
  31. Nesmeyanov, A.N.; Sazonova, V.A.; Drosd, V.N. Synthesis of Ferrocene Derivatives by Means of Boron- and Halogen-Substituted Ferrocenes. Chem. Ber. 1960, 93, 2717–2729. [Google Scholar]
  32. Abe, K.; Miyoshi, K.; Hattori, T.; Ushigoe, Y.; Yoshitake, H. Functional Electrolytes: Synergetic Effect of Electrolyte Additives for Lithium-Ion Battery. J. Power Sources 2008, 184, 449–455. [Google Scholar] [CrossRef]
  33. Chen, Y.; He, Q.; Mo, Y.; Zhou, W.; Zhao, Y.; Piao, N.; Liu, C.; Xiao, P.; Liu, H.; Li, B.; et al. Engineering an Insoluble Cathode Electrolyte Interphase Enabling High Performance NCM811//Graphite Pouch Cell at 60 °C. Adv. Energy Mater. 2022, 12, 2201631. [Google Scholar] [CrossRef]
  34. Kubota, T.; Ihara, M.; Katayama, S.; Nakai, H.; Ichikawa, J. 1,1-Difluoro-1-Alkenes as New Electrolyte Additives for Lithium Ion Batteries. J. Power Sources 2012, 207, 141–149. [Google Scholar] [CrossRef]
  35. Husch, T.; Yilmazer, N.D.; Balducci, A.; Korth, M. Large-Scale Virtual High-Throughput Screening for the Identification of New Battery Electrolyte Solvents: Computing Infrastructure and Collective Properties. Phys. Chem. Chem. Phys. 2015, 17, 3394–3401. [Google Scholar] [CrossRef]
  36. Abboud, J.-L.M.; Notari, R. Critical Compilation of Scales of Solvent Parameters. Part I. Pure, Non-Hydrogen Bond Donor Solvents. Pure Appl. Chem. 1999, 71, 645–718. [Google Scholar] [CrossRef]
  37. Yaws, C.L. Thermophysical Properties of Chemicals and Hydrocarbons; William Andrew: Norwich, NY, USA, 2008. [Google Scholar]
  38. Marcus, Y.; Kamlet, M.J.; Taft, R.W. Linear Solvation Energy Relationships: Standard Molar Gibbs Free Energies and Enthalpies of Transfer of Ions from Water into Nonaqueous Solvents. J. Phys. Chem. 1988, 92, 3613–3622. [Google Scholar] [CrossRef]
  39. Hansch, C.; Leo, A.; Hoekman, D. Exploring QSAR:  Hydrophobic, Electronic, and Steric Constants; ACS professional reference book; American Chemical Society: Washington, DC, USA, 1995. [Google Scholar]
  40. Linstrom, P.; Mallard, W.G. (Eds.) NIST Chemistry WebBook, NIST Standard Reference Database 69; National Institute of Standards and Testing (NIST): Gaithersburg, MD, USA, 1997. [Google Scholar] [CrossRef]
  41. Klein, J.; Khartabil, H.; Boisson, J.-C.; Contreras-García, J.; Piquemal, J.-P.; Hénon, E. New Way for Probing Bond Strength. J. Phys. Chem. A 2020, 124, 1850–1860. [Google Scholar] [CrossRef]
  42. Wiberg, K.B. Application of the Pople-Santry-Segal CNDO Method to the Cyclopropylcarbinyl and Cyclobutyl Cation and to Bicyclobutane. Tetrahedron 1968, 24, 1083–1096. [Google Scholar] [CrossRef]
  43. Clark, E. Sulfolane and Sulfones. In Kirk-Othmer Encyclopedia of Chemical Technology; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2000. [Google Scholar]
  44. Liu, J.; Shi, W.; Wang, X. ZnO–POM Cluster Sub-1 nm Nanosheets as Robust Catalysts for the Oxidation of Thioethers at Room Temperature. J. Am. Chem. Soc. 2021, 143, 16217–16225. [Google Scholar] [CrossRef]
  45. Bresciani, G.; Bortoluzzi, M.; Marchetti, F.; Pampaloni, G. Titanium(IV) Alkoxide-Carbamate Complexes: Synthesis and Catalytic Potential in H2O2-Oxidation of Organic Sulfides. Eur. J. Inorg. Chem. 2022, 2022, e202200402. [Google Scholar] [CrossRef]
  46. Assis, M.; Gouveia, A.F.; Ribeiro, L.K.; Ponce, M.A.; Churio, M.S.; Oliveira, O.N.; Mascaro, L.H.; Longo, E.; Llusar, R.; Guillamón, E.; et al. Towards an Efficient Selective Oxidation of Sulfides to Sulfones by NiWO4 and α-Ag2WO4. Appl. Catal. Gen. 2023, 652, 119038. [Google Scholar] [CrossRef]
  47. Kuźnik, N.; Krompiec, S.; Bieg, T.; Baj, S.; Skutil, K.; Chrobok, A. Double Bond Migration in S-allyl Systems Catalysed by [RuClH(CO)(PPh3)3]. J. Organomet. Chem. 2003, 665, 167–175. [Google Scholar] [CrossRef]
  48. Lim, K.M.-H.; Hayashi, T. Rhodium-Catalyzed Asymmetric Arylation of Allyl Sulfones under the Conditions of Isomerization into Alkenyl Sulfones. J. Am. Chem. Soc. 2015, 137, 3201–3204. [Google Scholar] [CrossRef] [PubMed]
  49. Jalilian, F.; Yadollahi, B.; Farsani, M.R.; Tangestaninejad, S.; Rudbari, H.A.; Habibi, R. New Perspective to Keplerate Polyoxomolybdates: Green Oxidation of Sulfides with Hydrogen Peroxide in Water. Catal. Commun. 2015, 66, 107–110. [Google Scholar] [CrossRef]
  50. Sasaki, Y.; Ue, M.; Takehara, M. Monofluorosulfolane. JP2002155074A, 28 May 2002. [Google Scholar]
  51. Sasaki, Y.; Ebara, R.; Nanbu, N.; Takehara, M.; Ue, M. Direct Fluorination of γ-Butyrolactone. J. Fluor. Chem. 2001, 108, 117–120. [Google Scholar] [CrossRef]
  52. Nielsen, M.K.; Ahneman, D.T.; Riera, O.; Doyle, A.G. Deoxyfluorination with Sulfonyl Fluorides: Navigating Reaction Space with Machine Learning. J. Am. Chem. Soc. 2018, 140, 5004–5008. [Google Scholar] [CrossRef]
  53. Deng, L.; Kang, B.; Englert, U.; Klankermayer, J.; Palkovits, R. Direct Hydrogenation of Biobased Carboxylic Acids Mediated by a Nitrogen-centered Tridentate Phosphine Ligand. ChemSusChem 2016, 9, 177–180. [Google Scholar] [CrossRef]
  54. Risley, J.M.; Kastanis, J.P.; Young, A.M. Determination of the Second-Order 1H NMR Parameters for the Aromatic Protons in 4-Fluoroaniline and Application to the Analysis of the 1H NMR Spectra for the Aromatic Protons in N4-(4′-Fluorophenyl)Succinamic Acid and in N4-(4′-Fluorophenyl)-3,3-Difluorosuccinamic Acid. J. Fluor. Chem. 2011, 132, 269–275. [Google Scholar] [CrossRef]
  55. Yu, Y.; Guo, Y.; Zhan, W.; Guo, Y.; Wang, Y.; Wang, Y.; Zhang, Z.; Lu, G. Gas-Phase Hydrogenation of Maleic Anhydride to γ-Butyrolactone at Atmospheric Pressure over Cu–CeO2–Al2O3 Catalyst. J. Mol. Catal. Chem. 2011, 337, 77–81. [Google Scholar] [CrossRef]
  56. Kobayashi, M.; Inoguchi, T.; Iida, T.; Tanioka, T.; Kumase, H.; Fukai, Y. Development of Direct Fluorination Technology for Application to Materials for Lithium Battery. J. Fluor. Chem. 2003, 120, 105–110. [Google Scholar] [CrossRef]
  57. Kvíala, J.; Plocar, J.; Vlasáková, R.; Paleta, O.; Pelter, A. 2-Fluoro-2-Buten-4-Olide, a New Fluorinated Synthon. Preparation; 1,2-, 1,4- and Tandem Additions. Synlett 1997, 1997, 986–988. [Google Scholar] [CrossRef]
  58. Ahn, J.A.; Jung, H.M.; Park, J.H.; Lee, C.H.; Lim, Y.M.; Ha, Y.J. Non-Aqueous Electrolyte and Secondary Battery Comprising the Same. KR20080065561A, 14 July 2008. [Google Scholar]
  59. Vasileff, T.P.; Koster, D.F. Preparation of Maleoyl Fluoride. Nuclear Magnetic Resonance Spectra of Maleoyl Fluoride, Fumaroyl Fluoride, and Fumaroyl Chloride Fluoride. J. Org. Chem. 1970, 35, 2461–2463. [Google Scholar] [CrossRef]
  60. Cao, R.; Liu, C.; Liu, L. A Convenient Synthesis of 2(5h)-Furanone. Org. Prep. Proced. Int. 1996, 28, 215–216. [Google Scholar] [CrossRef]
  61. Stavber, S.; Šket, B.; Zajc, B.; Zupan, M. Chemistry of Organo Halogenic Molecules. Part 100. Comparative Behaviour of Xenon Diflouride and Caesium Fluoroxysulphate in the Fluorination of Enol Acetates and Ketones. Tetrahedron 1989, 45, 6003–6010. [Google Scholar] [CrossRef]
  62. Stavber, S.; Zupan, M. High Yield Direct Fluorofunctionalisation of Ketones Using AccufluorTM-NFTh Fluorinating Reagent. Tetrahedron Lett. 1996, 37, 3591–3594. [Google Scholar] [CrossRef]
  63. Zajc, B.; Zupan, M. Room Temperature Fluorination of 1,3-Diketones and Enol Acetates with Xenon Difluoride. J. Chem. Soc. Chem. Commun. 1980, 759–760. [Google Scholar] [CrossRef]
  64. Weng, Z.; Wang, F. Salt and Crystal Form of Ketohexokinase Inhibitor and Use Thereof. WO2022022476A1, 3 February 2022. [Google Scholar]
  65. Kaneko, C.; Sato, M. Production of Alpha-Fluorocycloalkenones. JPH07242595A, 19 September 1995. [Google Scholar]
  66. Gharagheizi, F.; Eslamimanesh, A.; Mohammadi, A.H.; Richon, D. Empirical Method for Representing the Flash-Point Temperature of Pure Compounds. Ind. Eng. Chem. Res. 2011, 50, 5877–5880. [Google Scholar] [CrossRef]
  67. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 16, Revision A.03; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  68. Marenich, A.V.; Cramer, C.J.; Truhlar, D.G. Universal Solvation Model Based on Solute Electron Density and on a Continuum Model of the Solvent Defined by the Bulk Dielectric Constant and Atomic Surface Tensions. J. Phys. Chem. B 2009, 113, 6378–6396. [Google Scholar] [CrossRef]
  69. Lu, T.; Chen, F. Multiwfn: A Multifunctional Wavefunction Analyzer. J. Comput. Chem. 2012, 33, 580–592. [Google Scholar] [CrossRef]
  70. Curtiss, L.A.; Redfern, P.C.; Raghavachari, K. Gaussian-4 Theory Using Reduced Order Perturbation Theory. J. Chem. Phys. 2007, 127, 124105. [Google Scholar] [CrossRef] [PubMed]
  71. BIOVIA COSMOtherm, Release 2023; Dassault Systèmes. Available online: http://www.3ds.com (accessed on 22 November 2023).
  72. TURBOMOLE V6.5 2013, a Development of University of Karlsruhe and Forschungszentrum Karlsruhe GmbH, 1989–2007, TURBOMOLE GmbH, since 2007. Available online: http://www.turbomole.com (accessed on 22 November 2023).
  73. Klamt, A. Conductor-like Screening Model for Real Solvents: A New Approach to the Quantitative Calculation of Solvation Phenomena. J. Phys. Chem. 1995, 99, 2224–2235. [Google Scholar] [CrossRef]
  74. Klamt, A.; Jonas, V.; Bürger, T.; Lohrenz, J.C.W. Refinement and Parametrization of COSMO-RS. J. Phys. Chem. A 1998, 102, 5074–5085. [Google Scholar] [CrossRef]
  75. Eckert, F.; Klamt, A. Fast Solvent Screening via Quantum Chemistry: COSMO-RS Approach. AIChE J. 2002, 48, 369–385. [Google Scholar] [CrossRef]
  76. Klamt, A.; Eckert, F.; Arlt, W. COSMO-RS: An Alternative to Simulation for Calculating Thermodynamic Properties of Liquid Mixtures. Annu. Rev. Chem. Biomol. Eng. 2010, 1, 101–122. [Google Scholar] [CrossRef] [PubMed]
  77. Marenich, A.V.; Ho, J.; Coote, M.L.; Cramer, C.J.; Truhlar, D.G. Computational Electrochemistry: Prediction of Liquid-Phase Reduction Potentials. Phys. Chem. Chem. Phys. 2014, 16, 15068–15106. [Google Scholar] [CrossRef] [PubMed]
  78. Korth, M. Large-Scale Virtual High-Throughput Screening for the Identification of New Battery Electrolyte Solvents: Evaluation of Electronic Structure Theory Methods. Phys. Chem. Chem. Phys. 2014, 16, 7919–7926. [Google Scholar] [CrossRef]
  79. Tshepelevitsh, S.; Hernits, K.; Leito, I. Prediction of Partition and Distribution Coefficients in Various Solvent Pairs with COSMO-RS. J. Comput. Aided Mol. Des. 2018, 32, 711–722. [Google Scholar] [CrossRef]
  80. Dupeux, T.; Gaudin, T.; Marteau-Roussy, C.; Aubry, J.-M.; Nardello-Rataj, V. COSMO-RS as an Effective Tool for Predicting the Physicochemical Properties of Fragrance Raw Materials. Flavour Fragr. J. 2022, 37, 106–120. [Google Scholar] [CrossRef]
  81. Klamt, A.; Eckert, F.; Hornig, M.; Beck, M.E.; Bürger, T. Prediction of Aqueous Solubility of Drugs and Pesticides with COSMO-RS. J. Comput. Chem. 2002, 23, 275–281. [Google Scholar] [CrossRef]
  82. Kholod, Y.A.; Muratov, E.N.; Gorb, L.G.; Hill, F.C.; Artemenko, A.G.; Kuz’min, V.E.; Qasim, M.; Leszczynski, J. Application of Quantum Chemical Approximations to Environmental Problems: Prediction of Water Solubility for Nitro Compounds. Environ. Sci. Technol. 2009, 43, 9208–9215. [Google Scholar] [CrossRef] [PubMed]
  83. Chang, H.; Zhong, F.; Cao, R.; Liu, Y.; Wang, M.; Pu, X. Prediction of the Solubilities of Water in Hydrocarbons with COSMO-RS and Interpretation of the Solubility Characteristics. J. Solut. Chem. 2020, 49, 365–382. [Google Scholar] [CrossRef]
  84. Reinisch, J.; Klamt, A. Predicting Flash Points of Pure Compounds and Mixtures with COSMO-RS. Ind. Eng. Chem. Res. 2015, 54, 12974–12980. [Google Scholar] [CrossRef]
  85. University of Tartu. UT Rocket; share.neic.no. Available online: https://doi.org/10.23673/PH6N-0144 (accessed on 22 November 2023).
Figure 1. The studied molecules. The compound codes denote the parent structure (SL, GBL, or CP), number of double bond(s) (A—none, B/C—one, and D—two), and number of fluorine atoms. (a) Sulfolane derivatives, (b) cyclopentanone derivatives, and (c) GBL derivatives.
Figure 1. The studied molecules. The compound codes denote the parent structure (SL, GBL, or CP), number of double bond(s) (A—none, B/C—one, and D—two), and number of fluorine atoms. (a) Sulfolane derivatives, (b) cyclopentanone derivatives, and (c) GBL derivatives.
Molecules 28 07770 g001
Figure 2. Effect of fluorination on bond strength and multiplicity in SL derivatives. The plot is split by structural subgroups (AD), and double bonds are marked with wide vertical lines. The compound SL-B1b (mono-fluorinated at C5) is marked by a dashed line. The parameters of the S=O bonds 1-6 and 1-7 were averaged.
Figure 2. Effect of fluorination on bond strength and multiplicity in SL derivatives. The plot is split by structural subgroups (AD), and double bonds are marked with wide vertical lines. The compound SL-B1b (mono-fluorinated at C5) is marked by a dashed line. The parameters of the S=O bonds 1-6 and 1-7 were averaged.
Molecules 28 07770 g002
Figure 3. Effect of fluorination on bond strength and multiplicity in CP derivatives. The plot is split by structural subgroups (AD), and double bonds are marked with wide vertical lines. The compound CP-B1b (mono-fluorinated at C5) is marked by a dashed line.
Figure 3. Effect of fluorination on bond strength and multiplicity in CP derivatives. The plot is split by structural subgroups (AD), and double bonds are marked with wide vertical lines. The compound CP-B1b (mono-fluorinated at C5) is marked by a dashed line.
Molecules 28 07770 g003
Figure 4. Effect of fluorination on bond strength and multiplicity in GBL derivatives. The plot is split by structural subgroups (AC), and double bonds are marked with wide vertical lines. Solid lines denote the compounds fluorinated at C2 and dashed lines those at C4.
Figure 4. Effect of fluorination on bond strength and multiplicity in GBL derivatives. The plot is split by structural subgroups (AC), and double bonds are marked with wide vertical lines. Solid lines denote the compounds fluorinated at C2 and dashed lines those at C4.
Molecules 28 07770 g004
Figure 5. Computed oxidation and reduction free energies in the gas phase and DMSO relative to sulfolane (SL-A0). Black lines—gas phase; blue lines—DMSO. The values are not shown if, in any of the involved calculations, the structure broke apart or was rearranged during geometry optimization.
Figure 5. Computed oxidation and reduction free energies in the gas phase and DMSO relative to sulfolane (SL-A0). Black lines—gas phase; blue lines—DMSO. The values are not shown if, in any of the involved calculations, the structure broke apart or was rearranged during geometry optimization.
Molecules 28 07770 g005
Figure 6. HOMO and LUMO energies computed at the B3LYP/6-311+G** level of theory. Blue lines—LUMO levels; green lines—HOMO levels.
Figure 6. HOMO and LUMO energies computed at the B3LYP/6-311+G** level of theory. Blue lines—LUMO levels; green lines—HOMO levels.
Molecules 28 07770 g006
Figure 7. Effect of fluorination on compound properties. On the x-axis are the structural subgroups (A—no double bonds, B/C—one double bond, and D—two double bonds).
Figure 7. Effect of fluorination on compound properties. On the x-axis are the structural subgroups (A—no double bonds, B/C—one double bond, and D—two double bonds).
Molecules 28 07770 g007
Figure 8. Calculated solubility estimates (in mole fraction) and their ratios on the assumption that the solvent is liquid. No values are provided if miscibility was predicted.
Figure 8. Calculated solubility estimates (in mole fraction) and their ratios on the assumption that the solvent is liquid. No values are provided if miscibility was predicted.
Molecules 28 07770 g008
Figure 9. Normalized values of the essential properties of the studied compounds (for each property, the values of all represented compounds were rescaled to the range from 0 to 1).
Figure 9. Normalized values of the essential properties of the studied compounds (for each property, the values of all represented compounds were rescaled to the range from 0 to 1).
Molecules 28 07770 g009
Figure 10. Relationship between the experimental and computed ∆trG(H+) values. The dashed line is the regression line for the aprotic solvents protonating at the O atom. The numerical data are provided in Table S1.
Figure 10. Relationship between the experimental and computed ∆trG(H+) values. The dashed line is the regression line for the aprotic solvents protonating at the O atom. The numerical data are provided in Table S1.
Molecules 28 07770 g010
Figure 11. Correlation between the experimental and computational BP values. Fluorinated solvents are marked with squares.
Figure 11. Correlation between the experimental and computational BP values. Fluorinated solvents are marked with squares.
Molecules 28 07770 g011
Table 1. Literature data on the melting points (°C) of some of the studied compounds.
Table 1. Literature data on the melting points (°C) of some of the studied compounds.
Double BondsCompoundMPCompoundMPCompoundMP
0SL-A028.4 [24]CP-A0−52.8 [25]GBL-A0−43.1 [26]
1SL-B049–50 [27] GBL-B04–5 [28]
1SL-C063–64 [29] GBL-C048–49 [30]
2 CP-D096–98 [31]
Table 2. Estimated properties of the investigated molecules. ∆oxG and ∆redG values are relative to sulfolane (SL-A0). Where available, experimental values are provided with reference to the source. Three conventional solvents were added for comparison (MeCN—acetonitrile, DMSO—dimethyl sulfoxide, PC—propylene carbonate). Further computed data (pKaH(MeCN), GB, and solubility estimates) are available in Table S4.
Table 2. Estimated properties of the investigated molecules. ∆oxG and ∆redG values are relative to sulfolane (SL-A0). Where available, experimental values are provided with reference to the source. Three conventional solvents were added for comparison (MeCN—acetonitrile, DMSO—dimethyl sulfoxide, PC—propylene carbonate). Further computed data (pKaH(MeCN), GB, and solubility estimates) are available in Table S4.
CompoundCASRelative Red./Ox. Energies (kcal mol−1)EHOMO
(eV)
ELUMO
(eV)
εrBP (K)trG°(H+)
(kJ mol−1)
logPo/w
oxGgasredGgasoxGDMSOredGDMSO
MeCN75-05-847.64.334.4−19.8−9.27−0.3235.94 [36]355 [37]46.4 [38]−0.34 [39]
DMSO67-68-5−28.712.7−29.6 −6.49−0.1646.71 [36]464 [37]−19.4 [38]−1.35 [39]
PC108-32-711.52.32.3−14.4−8.36−0.3162.93 [36]515 [37]50 [38]−0.41 [39]
SL-A0126-33-00000−7.89−0.3642.13 [36]560 [37]44−0.77 [39]
SL-A1397248-09-8−2.0−0.8−2.6 −8.07−0.424758162−0.91
SL-A22413977-86-1−2.0−2.8 −8.21−0.525459674−0.44
SL-A32413977-87-21.1−2.8 −8.38−0.5243567920.19
SL-B01192-16-11.5−16.7−2.6−29.6−7.92−1.255960649−1.42
SL-B1a---3.9−21.6−3.9−36.6−8.02−1.226160165−0.55
SL-B1b2851432-77-24.3−23.45.3−35.0−8.27−1.605259767−0.93
SL-B2---4.4−29.16.3−39.0−8.45−1.825860677−0.28
SL-B3---10.9−31.64.8−42.1−8.56−1.7850569970.78
SL-C077-79-21.0−6.9−13.8−19.5−8.02−0.834055356−0.90
SL-C1444334-21-8 −23.3 −8.22−1.634356572−0.68
SL-C2--- −27.7 −38.0−8.41−1.815258384−0.03
SL-C3--- −40.8 −50.7−8.59−2.39425551000.59
SL-D027092-46-2−1.9−46.6−9.4−60.6−7.66−2.864656870−0.63
SL-D1---−3.9−49.4−10.2−62.8−7.66−2.8946545860.34
CP-A0120-92-3−18.33.5−19.7−17.0−6.83−0.8614.45 [36]404 [37]120.45
CP-A11755-12-0−14.7 −15.5 −7.20−1.6329457380.10
CP-A22167972-33-8−11.0−20.9−10.4−40.3−7.48−1.8731456650.88
CP-B0930-30-3−16.3−15.0−16.3−34.2−6.90−1.6533409 [40]−4−0.11
CP-B1a143998-28-1−3.6−21.6−11.9−38.9−7.31−1.9440484160.18
CP-B1b---−11.6−26.7−12.2−44.6−7.25−2.194349922−0.30
CP-B2---−7.2−34.8−5.3−51.1−7.56−2.5746499440.40
CP-C014320-37-7−14.83.4−25.9−18.9−7.02−0.9614399250.55
CP-C1175544-12-4 (R)−10.0 −16.2 −7.37−1.7428450570.32
CP-C2---−5.1−25.7−5.3−45.1−7.67−2.1433453800.94
CP-D013177-38-3−16.6−48.0−24.8−69.1−7.04−3.1622397470.86
CP-D1---−17.6−53.8−25.3−72.9−7.03−3.4027403661.22
GBL-A096-48-00.95.62.0−14.4−7.66−0.2340.96 [36]477 [37]22−0.64 [39]
GBL-A1a3885-31-25.62.26.4 −7.99−1.114550446−0.39
GBL-A1b2343-90-010.24.310.0−21.2−8.16−0.673247045−0.13
GBL-A2a220294-13-39.4−13.711.2 −8.27−1.2644489640.33
GBL-A2b1345047-11-117.8 17.5 −8.58−1.0027452610.63
GBL-B0497-23-43.9−17.1−0.3−34.2−7.87−1.765651421−0.70
GBL-B1a197096-95-011.6−21.5−0.2 −8.24−1.925551140−0.12
GBL-B1b1052601-43-016.3−34.713.5−50.8−8.46−2.5134464520.22
GBL-B224647-21-038.3−44.122.4−58.7−8.92−2.9025418771.17
GBL-C020825-71-2−17.90.9−26.4−19.4−7.19−0.9431437410.11
GBL-C1a---−29.6 −27.6 −7.83−2.0330432760.60
GBL-C1b1052601-43-0−16.6 −24.8 −7.31−1.1222408610.88
GBL-C2---5.8−32.9−2.6−52.1−8.29−2.44294041031.42
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Tshepelevitsh, S.; Kütt, A.; Leito, I. Theoretical Study on Fluorinated Derivatives of Sulfolane, Cyclopentanone, and Gamma-Butyrolactone. Molecules 2023, 28, 7770. https://doi.org/10.3390/molecules28237770

AMA Style

Tshepelevitsh S, Kütt A, Leito I. Theoretical Study on Fluorinated Derivatives of Sulfolane, Cyclopentanone, and Gamma-Butyrolactone. Molecules. 2023; 28(23):7770. https://doi.org/10.3390/molecules28237770

Chicago/Turabian Style

Tshepelevitsh, Sofja, Agnes Kütt, and Ivo Leito. 2023. "Theoretical Study on Fluorinated Derivatives of Sulfolane, Cyclopentanone, and Gamma-Butyrolactone" Molecules 28, no. 23: 7770. https://doi.org/10.3390/molecules28237770

Article Metrics

Back to TopTop