Figure 1a shows the hierarchical down-selection screening approach we employed to identify new FLS for Li-S batteries. We began with a search of the solvents reported in Li-S literature and identified 14 FLS compounds (Supplementary Table S1). We performed a structure similarity search using these compounds as inputs to query the PubChem38 database and retrieved 1,250 fluorinated ether compounds. We filtered these compounds for neutral, non-aromatic, and C-O-H-F chemical systems, which reduced the size of the initial library to 922 compounds. At this stage, we performed path-based fingerprinting using OpenBabel39 and used the resulting fingerprints to calculate the structural similarity between each pair of molecules via the Tanimoto coefficient. The similarity scores were used to quantify the structural diversity of the library. A similarity score of zero indicates dissimilar structures, while a value of one indicates identical molecules. The cumulative distribution function (CDF) in Fig. 1b shows that only 10% of the pairs have a similarity score above 0.75. Given that the library is restricted to fluorinated ether molecules, this level of diversity is considered sufficient. Supplementary Fig. 1 shows the distribution of similarity scores between the compounds in the library and each of the 14 FLS reported in the literature. Most of the distributions are unimodal and indicate an intermediate degree of similarity with previously reported solvents. This is desirable because a high similarity score implies that the solvents closely resemble those previously tested, while a low similarity score suggests that the library has deviated from the ones reported in the literature with promising properties40.
In the next screening stage, we computed the solubility of lithium bis(trifluoromethane sulfonel) imide (LiTFSI) and lithium bis(fluorosulfonyl) imide (LiFSI) in each of the FLS using COSMO-RS calculations. LiTFSI and LiFSI were selected for this study due to their widespread application in Li-S batteries17. An ideal FLS should be capable of suppressing polysulfide solubility and dissolving sufficient amount of Li salt (~ 1 M) to facilitate the fast Li+ ion transport between the electrode while maintaining a low viscosity to enhance the wetting ability of the electrolyte to cell components such as electrode and separators. Out of the 922 solvents, only 141 compounds met the criterion of having Li salt solubility ≥ 1 M. All the FLS exhibited lower salt solubilities than conventional dioxolane (DOL) and 1,2-dimethoxyethane (DME) solvents. This result is expected since the solvating power of FLS molecules is restrained by the inductive effect imposed by the electron-withdrawing fluoroalkyl groups41. The solvation ability of FLS is not, however, necessarily inversely proportional to their fluorination degree, as will be explained later.
Safety considerations were accounted for by estimating the boiling point (BP) and flash point (FP) of the pure molecules using COSMO-RS calculations. BP and FP of electrolyte are critical parameters to ensure safe battery operation and reduce the potential harm to the batteries in the event of thermal runaway or ignition. Out of the 141 compounds, 86 met the criteria of BP ≥ 47 ℃ and FP ≥ -23 ℃. These parameters were selected to align with the acceptable and optimal temperature ranges for Li-ion batteries, which are − 20 ℃ to 60 ℃ and 15 ℃ to 35 ℃, respectively42. The BP threshold of 47 ℃ corresponds to the upper limit set by the majority of battery thermal management systems43. Supplementary Table S2 shows the predicted and measured BP for DME and common FLS. The mean absolute deviation (MAD, 16 ℃) in the order of ~ 20% of the BP window and the Pearson’s R value of 0.93 indicate a sufficient level of correlation.
Ionization potential (IP) and electron affinity (EA) were computed from DFT calculations and used as proxies for the solvent stability towards initial oxidation and reduction, respectively. These properties were then used to determine the electrochemical window of the solvent, within which the reversible reactions of the anode and cathode are facilitated without the continuous decomposition of the electrolyte components. Considering that the working voltage of sulfur cathode is between 1.5 and 3.0 V vs. Li/Li+ in Li-S batteries44,45, solvents with an electrochemical window greater than or equal to 2.7 V were down-selected to ensure electrochemical stability. Only two solvents from the previous step did not meet this criterion. Additional DFT calculations were performed to compute the relative binding energy (BE) of FLS with both LiTFSI and lithium polysulfide (LiPS), modeled as Li2S4. BE of DME with LiTFSI and Li2S4, -24.7 and − 22.3 kcal/mol, respectively, were used as reference zero. BE is calculated at the FLS ether oxygen-Li+ site for both FLS-LiTFSI and FLS-LiPS interactions. A desirable property of a potential FLS is higher selective solvating power than DME (second quadrant of Fig. 2a), i.e., the solvent is characterized with similar or higher BE with the weakly coordinating salt than DME but less binding affinity towards LiPS. As shown in Fig. 1a, 18 solvents out of 84 compounds selected from the previous step met this criterion. We note that DFT calculations were performed for the entire library of 922 compounds, extending beyond the subset identified through screening at that stage (Refer to Supplementary Fig. 2 for the distribution of computed properties of FLS in comparison with DOL, DME, and fluorinated ether solvents reported in Li-S literature). Figure 2b shows that fluorination degree has more significant positive impact on the solvent IP than its BE with LiTFSI. Here, fluorination degree refers to the percentage of hydrogen atoms substituted by fluorine atoms in a molecule. While many solvents have the same fluorination degree, their ionization potentials and salt binding affinities can differ as shown in Fig. 2b.
To understand these correlations, we employed an extreme gradient boosting (XGBoost) regression model that allows identifying attributes of high impact on FLS properties relevant to Li-S electrolytes. We featurized the FLS molecules based on their structural properties, such as shape index, steric hinderance, Verloop’s Sterimol parameters46, radius, proximity of fluorine to oxygen, fluorination degree, among others. We also incorporated DFT-extracted properties like dipole moment, polarizability, and electrostatic partial charges (minimum and maximum absolute partial charges in the molecule). Refer to Supplementary Table S3 for detailed feature descriptions. We then used a Shapley Additive exPlanations (SHAP) framework47 to interpret predictions made by the XGBoost model. This technique utilizes Shapley values, a game theory-derived metric that assesses the contribution by different players. In this case, the players are the derived FLS features, and the game outcome is the model predictions of IP and LiTFSI BE. Figure 2c presents a summary plot that demonstrates the effect of the top 10 most essential features on the solvent IP, ordered according to their importance. A similar analysis for EA and BE is shown in Supplementary Fig. 3. The IP is primarily governed by the fluorination degree in a positive relation, as discussed earlier. The proximity of fluorine to the ether oxygen atom is the second most critical factor, exhibiting an overall negative impact on IP. Steric effect also significantly influences IP, with increased steric hindrance on the oxygen atom leading to a higher IP.
SHAP analysis for EA of FLS molecules (Supplementary Fig. 3a) indicates that fluorination degree is also the most impactful feature on reduction. However, steric effects, unlike oxidation, are negligible under reduction, similar to previously reported results for phenothiazines48. This behavior can be rationalized by the fact that the oxygen atom contributes considerably to the highest occupied molecular orbital but not much to the lowest unoccupied molecular orbital (Supplementary Fig. 4). Supplementary Fig. 3b shows that in general, a wider separation between fluorine and ether oxygens, a lower fluorination degree, and a higher FLS polarizability are correlated with an enhanced binding affinity of the solvent to LiTFSI. Supplementary Fig. 5 illustrates two molecules, both with a fluorination degree of 50%, yet with notably different IPs of 7.35 eV and 5.84 eV, and PS binding affinities of -16.5 kcal/mol and − 22.3 kcal/mol, respectively. A detailed examination reveals differing steric effects and fluorine-to-oxygen distances in these molecules: the left molecule shows a greater steric effect (2.41) and a closer fluorine atom, approximately 2.25 Å to the ether oxygen, in contrast to the corresponding values of 2.28 and 2.76 Å in the second molecule. These variations corroborate the SHAP analysis previously discussed, serving as an example of the discussed trends. Consequently, controlling factors like the fluorination degree, fluorine proximity to oxygen, and steric effects is crucial for manipulating IPs and binding affinities. Based on these observations, for a specific fluorination degree and to attain higher IPs and reduced PS solubility, we suggest adjusting the FLS structure to increase steric hindrance around the ether oxygen and position fluorine atoms closer to it.
For a thorough evaluation, we created two additional models, one that integrated 202 general RDKit49 descriptors and another with 1,323 Mordred50 descriptors, each combined with the custom FLS features described above. Supplementary Fig. 6 displays the SHAP analysis for IP, highlighting the top 10 features predicted from the decision tree model with a maximum tree depth of three to avoid overfitting. Upon incorporating RDKit and Mordred descriptors, the fluorination degree, proximity of fluorine to oxygen, and steric effect persist as the three dominant factors influencing IP. Among the top ten features, eight to nine were custom calculated FLS features. This finding emphasizes the significant advantage of developing features tailored to the specific nuances of the chemical system under study, as opposed to exclusive dependence on generic molecular descriptors. Moreover, the custom calculated FLS features offer an additional benefit in that they can be more readily fine-tuned experimentally, which is a flexibility often not afforded by many of the standardized RDKit or Mordred descriptors.
Supplementary Table S4 presents the structures of the 18 solvents selected from the last screening step. For each of these solvents, we performed MD simulations to gain insight into solvation and transport behavior. MD systems were composed of 1 M LiTFSI, 0.3 M LiNO3 in DOL/FLS (1/1; v/v) with and without the presence of 0.25 M Li2S2, 0.25 M Li2S4, and 0.25 M Li2S8, each run separately. FLS0 was included in these simulations despite not being screened, due to its commercial availability and its similar properties to the commonly used TTE solvent. We calculated radial distribution functions (RDF; Supplementary Fig. 7–11) and probability of forming a polysulfide of cluster sizes 1–6 in the first and second solvation shell of each PS specie (Fig. 3). Cluster analysis results indicate a lower solubility of Li2S2 compared to longer chain PS (Li2S4 and Li2S8) in all electrolytes including the state-of-the-art DOL/DME, in accordance with previously reported experimental results51. The introduction of FLS leads to a lower solubility of Li2S4 and Li2S8 (more clustering or aggregation), compared to the system containing DME, as indicated by lighter red and purple hues for PS size one, and darker tones for larger PS sizes (See sample MD snapshots in Supplementary Fig. 12–14). Even solvents with a fluorination degree less than 20% (FLS2, FLS4, FLS6, FLS10, FLS12, FLS14, FLS16, FLS17, and FLS18; see Supplementary Table S4) led to substantial PS aggregation, particularly in Li2S8 systems. This is an interesting result indicating the effectiveness of these solvents in reducing the solubility of long chain PS despite their low fluorination degree. Recent studies have increasingly shown that highly fluorinated solvents may not always be preferable, with several groups highlighting the need to explore partially fluorinated electrolytes52–54. We note that all screened FLS in this work, except FLS0, exhibit a lower fluorination degree than TTE.
The RDF analysis of \({\text{S}}_{\text{x}}^{2-}\)–solvent for various PS species (Supplementary Fig. 8) shows significantly weaker PS interactions with TTE and FLS than with DME. Such dramatic reduction in PS-solvent interaction in fluorinated systems explains the previously discussed long-range PS structures and consequently, the lower PS solubility in these systems. Additionally, Li-FLS interactions are weaker than the Li-DME counterpart (Supplementary Fig. 9), leading to more contact ion pairs in the FLS electrolytes, evidenced by the pronounced Li+-TFSI- RDF peaks at around 2 Å (Supplementary Fig. 10). Additionally, the less coordinating DOL solvent compared to DME, has more opportunity to interact with Li+ in these systems than in the DOL/DME electrolyte (Supplementary Fig. 11). However, improved Li+ solvation occurs in electrolytes containing FLS2, FLS12, and FLS14-FLS18 (peak at 1.9 A in Li+-solvent RDF; Supplementary Fig. 9), suggesting the potential of these solvents to reduce PS solubility while retaining a degree of coordinating ability. Despite both having a fluorination degree of around 17%, FLS4 and FLS12 exhibit distinct behaviors in terms of Li2S8 aggregation, with FLS4 leading to higher aggregation levels (Fig. 3). This difference can be attributed to the greater steric hindrance around the ether oxygen atom in FLS4 coupled with the closer proximity of fluorine atoms to oxygen compared to FLS12. These observations are in alignment with the SHAP trends previously discussed, emphasizing the influence of fluorine atom placement on FLS solvating power. Finally, we obtained the distribution of diffusion coefficients of Li+ and the co-solvent (DME, TTE, or FLS) within each of the electrolyte systems (Supplementary Fig. 15). Compared to the FLS-containing systems, much more significant overlap occurs between the diffusion coefficients of the cation and the co-solvent in the DOL/DME system, indicating more correlated motion compared to the FLS systems. This dynamical behavior reinforces the less coordinating capability of these solvents compared to DME.
One of the critical targets of this work is to improve the ionic conductivity and address transport challenges associated with fluorinated solvents, while retaining their effectiveness in mitigating the PS shuttle effect. To this end, we selected FLS0 and FLS18 for further evaluation due to their commercial availability. Figure 4a shows the ionic conductivity and viscosity of electrolytes containing DME, TTE, FLS0, and FLS18, used as cosolvents with DOL. Bars represent results from MD simulations, whereas stars indicate experimental data collected at 25 ℃. All four electrolytes containing the new FLS exhibit lower viscosities and superior ionic conductivities compared to their TTE-based counterpart, while maintaining viscosities (< 2 cP) on par with the state-of-the-art DOL/DME system. Notably, the two equivolume ternary mixtures of DOL/DME/FLS0 and DOL/DME/FLS18 achieve ionic conductivities of 8.6 and 9.9 mS/cm, respectively. These values are comparable to state-of-the-art DOL/DME system and surpass the ionic conductivities of similar ternary mixtures containing fluorinated solvents20,55,56 (note good agreement of MD predictions with measured viscosity, density, and ionic conductivity as function of temperature in Supplementary Fig. 16). Even binary mixtures of DOL/FLS show conductivities 1.5–2.3 times higher than that of DOL/TTE in the presence of 0.1 M LiNO3, which can be attributed to the formation of less compact cation-anion clusters within these systems.
Encouraged by their potential in reducing PS solubility and their improved dynamical properties over the widely used TTE, we performed Li || S/C full battery tests of the following four electrolytes: 1 M LiTFSI, 0.3 M LiNO3 in DOL/DME, 1 M LiTFSI, 0.3 M LiNO3 in DOL/TTE, 1 M LiTFSI, 0.3 M LiNO3 in DOL/DME/FLS0, and 1 M LiTFSI, 0.1 M LiNO3 in DOL/DME/FLS18, each with composite sulfur/carbon cathode at ~ 1 mg cm− 2 S loading and a glass fiber paper separator, with constant current cycling at C/10 rate. Here, it is important to note that we encountered solubility issues with LiNO3 in the fluorinated electrolytes. Specifically, a maximum concentration of 0.1 M LiNO3 could be dissolved in systems containing DOL/FLS0, DOL/FLS18, and DOL/DME/FLS18. In contrast, the DOL/DME/FLS0 system could dissolve up to 0.3 M LiNO3. This observation underlines a limitation in our computational screening, which initially focused only on the solubility of the Li salt and not the additive. Consequently, future developments in designing these fluorinated solvents should also consider the additives solubility.
Figure 4 shows representative charge and discharge curves along with the discharge capacity and Coulombic efficiency (CE) as a function of the cycle number. The capacity fade is alleviated upon the use of FLS0 and TTE relative to that of the base DOL/DME electrolyte, consistent with prior results using TTE as a cosolvent57. The degree of capacity retention between DOL/TTE and DOL/DME/FLS0 is comparable yielding similar discharge capacities at 200 cycles. Overall, the cell with DOL/DME/FLS0 attains the highest capacity after 200 cycles and the least capacity fading (Fig. 4b) followed by DOL/TTE. The cell with DOL/DME/FLS18 resulted in visible capacity degradation to ~ 9% within 200 cycles, likely attributed to LiNO3 solubility issues and the strong reaction of FLS18 with Li metal in the presence of the additive, thus complicating interphase formation. At cycle 200th, cells with DOL/DME/FLS0 and DOL/TTE exhibit discharge capacities of 243 and 240 mAh/g, respectively, whereas the DOL/DME and DOL/DME/FLS18 systems resulted in capacities of only 55 and 86 mAh/g, respectively. Li-S cells using TTE and FLS presented more stable cycling than that containing DOL/DME, indicating a reduction in the shuttling effect due to the decrease of LiPS solubility.
Figure 4b shows the discharge and charge profiles for the same Li-S cells at cycle 10. The voltage of the first plateau corresponding to the solid S conversion to LiPSs is extracted at 15 mAh/g, and the second plateau corresponding to the LiPSs conversion to Li2S is at 500 mAh/g. For the system containing DOL/DME/FLS0, the first plateau occurs at a lower voltage compared to the DOL/DME cell, while the second plateau occurs at a higher voltage (Supplementary Fig. 17). This pattern indicates that decreasing LiPS solubility suppresses the formation of LiPSs, thus reducing the shuttling effect. The system with DOL/DME/FLS18 showed high overpotential and low capacity utilization likely attributed to the solubility and reactivity issues mentioned earlier. Overall, the cell with DOL/DME/FLS0 exhibited the best performance with the highest discharge capacity of 785 mAh/g, compared to 708 mAh/g for the DOL/TTE system.
The lower CE of the DOL/DME/FLS0 cell compared to that of DOL/DME (Fig. 4d) indicates reduced self-discharge behavior due to the polysulfide shuttling effect. The dissolution of active S into the electrolyte as well as redox shuttling of soluble polysulfide chains during charging results in Li-S batteries often having a CE > 100%, where CE = Charge Capacity/Discharge Capacity58. This overcharge effect may be even further magnified at a slow charge/discharge rate such as C/10 used herein. Improved CEs through screened FLS are observed in the voltage profiles in Fig. 4c, which show less additional capacity in the charge profile. In Cycle 10, DOL/DME, DOL/TTE, DOL/DME/FLS0, and DOL/DME/FLS18 show CEs of 104.6, 103.1, 100.9, and 102.6%, respectively. Figure 4d further shows that the DOL/DME/FLS0 and DOL/DME/FLS18 electrolytes screened solvents can maintain reasonable CEs and mitigate the shuttling effect over a long duration of 200 cycles.
In summary, we demonstrate a systematic high-throughput screening approach to identify new FLS exhibiting a promising balance between transport properties and PS solubility, outperforming the widely used fluorinated solvent, TTE. Interestingly, we find that even solvents with a low fluorination degree can potentially mitigate the shuttle effect in Li-S batteries. Such a feature is increasingly desirable for its potential to enhance battery performance while maintaining environmental considerations. However, our findings also suggest a need for further optimization of the electrolyte composition to address solubility issues associated with other electrolyte additives such as LiNO3. We also highlight on three highly tunable properties of FLS—fluorination degree, oxygen steric effect, and proximity of fluorine atoms to ether oxygen—as critical features for refining solvent design. The SHAP analysis results emphasize the significance of developing tailored features that cater to the specific nuances of the chemical system under study, rather than depending on generic molecular descriptors. Although our evaluation was limited to only two solvents due to their commercial availability, we have revealed an additional 17 solvents with promising properties for future experimental investigation. This work informs and inspires viable paths forward for the future design and optimization of next-generation Li-S electrolyte solutions.