Next Article in Journal
Ordered Mesoporous Carbon as a Support of Pd Catalysts for CO2 Electrochemical Reduction
Next Article in Special Issue
A Hybrid Microbial–Enzymatic Fuel Cell Cathode Overcomes Enzyme Inactivation Limits in Biological Fuel Cells
Previous Article in Journal
A Review on Catalysts Development for Steam Reforming of Biodiesel Derived Glycerol; Promoters and Supports
Previous Article in Special Issue
Outlook on the Role of Microbial Fuel Cells in Remediation of Environmental Pollutants with Electricity Generation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

A DFT Investigation on the Origins of Solvent-Dependent Polysulfide Reduction Mechanism in Rechargeable Li-S Batteries

Department of Chemistry, National Taiwan Normal University No. 88, Section 4, Tingchow Road, Taipei 116, Taiwan
*
Author to whom correspondence should be addressed.
Catalysts 2020, 10(8), 911; https://doi.org/10.3390/catal10080911
Submission received: 14 May 2020 / Revised: 30 July 2020 / Accepted: 4 August 2020 / Published: 10 August 2020
(This article belongs to the Special Issue Catalysts for Microbial Fuel Cells)

Abstract

:
The lithium-sulfur (Li-S) battery is one of the promising energy storage alternatives because of its high theoretical capacity and energy density. Factors governing the stability of polysulfide intermediates in Li-S batteries are complex and are strongly affected by the solvent used. Herein, the polysulfide reduction and the bond cleavage reactions are calculated in different solvent environments by the density functional theory (DFT) methods. We investigate the relationship between the donor numbers (DN) as well as the dielectric constants (ε) of the solvent system and the relative stability of different polysulfide intermediates. Our results show that the polysulfide reduction mechanism is dominated by its tendency to form the ion-pair with Li+ in different organic solvents.

Graphical Abstract

1. Introduction

Lithium-sulfur (Li-S) batteries, with their high theoretical energy density of 2600 Wh/kg, are attracting much attention as future energy storage systems in various fields, including portable electronic devices, hybrid to full electric vehicles (EVs), and large-scale power grids [1,2]. Since the sulfur cathode could undergo multiple electron transfer events during the Li-S battery operations, a high theoretical capacity (1675 mAh/g) can be achieved. In comparison, the conventional intercalated LiCoO2 cathode in the lithium-ion batteries (LIB) shows a capacity of around only 140 mAh/g [3]. The abundant quantity and the low cost of the sulfur resources on earth also guarantee economic advantages of Li-S batteries [4,5,6]. Overall, the predicted high efficiency and the environmental compatibility of the Li-S batteries have made them the most promising candidate to overcome the intrinsic limitation of the common batteries for the next generation rechargeable energy storage system [4,7,8,9].
However, the development on the application of Li-S batteries is hindered by various problems, including the shuttle effect [9,10,11,12] and the dendrite formation at the anode [13,14], etc. The combined effects of the physical properties, such as viscosity, conductivity, and safety issues of the electrolyte system [15,16], the salt additives [17], electrode materials [18,19,20,21], interface stabilities [22,23,24,25], etc., have been the subject of intense investigations in the past decade aiming for the practical and sustainable scaling of high-performance Li-S batteries [26,27]. The physical environment of a Li-S battery is an extremely complicated system and the plethora of experimental attempts are still largely based on trial and error. Large-scale theoretical simulations have also been carried out to improve the fundamental understanding of the structural and transport behaviors of the intermediates, the interaction between the electrolyte and the reacting species, and the adsorption event at the electrode surfaces [28,29,30,31,32,33].
In principle, the selection of the solvent and the electrolyte system is critical to the battery performance since the sulfur reduction mechanisms [10,34,35,36] depend strongly on the dissolution, diffusion and disproportionation reactions of soluble polysulfides in the solvent environment [5,9,37,38,39,40,41,42,43,44]. The chain length distribution of the polysulfide species intimately affects the diffusion coefficients, the conductivities, clustering, and even the extent of the shuttle effect [12,28,29]. Meanwhile, the major polysulfide species may exist in different forms [44] depending on the solvent system [38,43,45], the ionic additives [46], and even the concentrations [43]. For example, a previous study has shown that different alkali-metal cations could stabilize polysulfides of different length, e.g., S42− or S32− [38]. It was also observed that the radical polysulfide species are favored at lower sulfur concentrations, while the doubly charged, dimerized polysulfide species are favored at higher sulfur concentrations [43] in accordance with chemical equilibrium. For the more fundamental solvent effect, different theories have been suggested to rationalize experimental observations. For example, the most stable polysulfide intermediate may exist as S3 radicals in dimethyl sulfoxide (DMSO), or as S42− in the mixed solvent system, 1,3-dioxane:1,2-dimethoxyethane (DOL:DME) [38,47]. A different study reported that polysulfide radicals (i.e., S4) can be selectively stabilized in ether-based electrolytes, such as in tetraethylene glycol dimethyl ether (TEGDME), and polyethylene oxide (PEO) [43,48]. These phenomena were tentatively correlated with the dielectric constant (ε) and/or the donor number (DN) of the solvent system. Nevertheless, a microscopic understanding is lacking. In general, the ε of the solvent is related to the stability of polar or ionic solute species, and the DN is directly related to the complexation between the cation and the solvent molecules. Therefore, a single parameter may not offer a comprehensive explanation to the observed tendency of the dominant polysulfide species. A more profound understanding of the multiple polysulfide reduction pathways in different electrolytes is necessary.
In this work, we focus on the interaction between polysulfides, lithium ions, and the solvent system. The structures of the polysulfides and the lithium polysulfides (LiPSs) in different solvent systems are simulated, and various possible reduction pathways of these species are characterized. We consider both the implicit solvent model and the explicit solvent binding with Li+. Our results shall help establish the dominant factors of polysulfide stabilities, which may be applied to solvent system optimization in Li-S battery design.

2. Results and Discussions

The optimized geometries of polysulfides (PS) and lithium polysulfides (LiPS) in DMSO are shown in Supplementary Figure S1 and Figure 1, respectively. Note that the overall geometries are similar in different solvents. The lengths of S-S bonds are in the range of 1.99 to 2.19 Å for both PS’s and LiPS’s, and the lengths of Li-S bonds are in the range of 2.37 to 2.44 Å. The overall geometry and the detailed geometry parameters are consistent with previously reported values [29,33,49,50,51]. We then consider the electrochemistry profile and compute the reduction potentials for major polysulfide reduction reactions, as shown in Supplementary Table S1. Note that since the mixed-solvent system DOL:DME cannot be easily simulated by the implicit solvent model, we adopt the THF solvent system which has a similar molecular motif and dielectric constant (εTHF = 7.43) value with DOL and DME (εDOL = 7.3 and εDME = 7.2). As can be expected, the calculated reduction potentials increase in solvents with higher dielectric constant as the anions are stabilized in polar solvents. Overall, the computed theoretical values are consistent with the cyclic voltammogram experimental values.

2.1. Solvent-Dependent PS Reaction Pathways

To further explore the reactions of PS in different solvents, we consider some possible reaction pathways (Reactions (1)–(8) in Table 1) based on the results from UV-Vis spectroscopy [38,52,53], and we calculate their reaction free energies (∆G). In solvents with medium to low polarity (ε < 10), it has been proposed that the PS species may exist as neutral lithium polysulfides (LiPS) [44]. Therefore, the aforementioned reaction pathways between the undissociated LiPS reactants are also considered as Reactions (9)–(16) in Table 2.
The absorption spectrum of the PS’s in DMSO indicates the existence of the species S82, S62, S42−, and S3 [11]. In contrast, the reactions between sulfur and various lithium polysulfides in DOL:DME (represented by THF in our theoretical simulation) mainly produce the species S82 and S42− [38]. Therefore, for the initial S82 decomposition step, several crucial reactions are considered as shown by Reactions (1)–(4). Although the neutral cyclic S5 molecules and the radical anions S5 have not been experimentally observed in solution or in the solid state [44], their existence cannot be completely ruled out and are therefore considered as the decomposition product within the first few steps (Reaction (2)). Further decompositions of the short-chain PS species (S62−) are proposed to proceed by Reactions (5) and (6). The origins of the observed species may also come from disproportionation reactions such as Reactions (7) and (8). In general, for the same reaction, the reaction Gibbs free energy may monotonously decrease (Reactions (1) and (3)), or increase (Reactions (2) and (4)) with increasing solvent dielectric constant (ε) from DIOX to DMSO. Reaction (3) may be ruled out based on its high endothermicity (∆G > 1.10 eV) in all solvent systems.
For solvents with a high dielectric constant, the polysulfide intermediates are expected to exist as dissociated ionic species. For example, in the DMSO, the reaction energies of three considered decomposition reactions of S82− (Reactions (1), (2), and (4)) are all slightly endothermic (∆G ~ 0.3 eV), and are all likely to proceed. The existence of the decomposed products S62 (Reaction (1)) and S3 (Reaction (2)) have been observed experimentally [38]. The S62 may also easily undergo a bond cleavage to form two S3 radical anions, as shown by the slightly exothermic Reaction (5). In comparison, the decomposition of S62 to form S42 (Reaction (6)) involves a higher reaction energy (0.62 eV) and its occurrence is deemed more unlikely. The experimentally [11] observed S42 may result from the reduction of the S4 product in Reaction (4). Similarly, the relatively higher reaction Gibbs free energies involved in the disproportionation reactions in Reactions (7) and (8) have rendered these pathways rather improbable.
Unlike the reactions between the ionic species as in Table 1, the ∆G’s for neutral reactants are much less dependent on the ε, as the numerical values of the ∆G’s for the same reaction are all relatively similar in different solvents in Table 2. In solvents with low ε, e.g., DIOX or THF, the polysulfide intermediates are expected to exist as the neutral bound LiPS species. For these systems, the results indicate that the more probable decomposition pathways are Reaction (9) followed by Reaction (14), in which the Li2S8 species first decompose into Li2S6, and then to Li2S4. Recall that the major species observed in DOL:DME are S82− and S42 [38], which is consistent with our calculation results. Other reaction pathways generating the LiS3 (Reactions (10), (13), and (15) in Table 2) are deemed less probable due to their high endothermic characters (∆G > 0.5 eV).

2.2. The Effect of Solvent Donor Number (DN)

In a few cases, it has been observed that a high dielectric solvent may also facilitate the reaction pathway of the LiPS rather than the dissociated PS [38]. It has been proposed that, other than the dielectric constant, the donor number (DN) is another important factor governing solvent reactions. The donor number, defined as the negative enthalpy value for the 1:1 adduct formation in the dilute solution of a Lewis base and the standard Lewis acid antimony pentachloride (SbCl5) in the non-coordinating solvent 1,2-dichloroethane [54], is a parameter describing the cation coordinating power of solvents. To quantify the effect of the DN, we calculate the average binding energies ( E b ¯ ) for three high dielectric solvent systems, with moderate to high donor numbers. The estimated coordination numbers and the optimized geometries are shown in Table 3 and Supplementary Figure S2, respectively. The cutoff binding energy for the predicted maximal coordination number (np) is set as −0.30 eV based on the experimental observation of [Li-(DMSO)4]+ [46]. As can be expected, the value of np increases with the DN. The calculated small coordination number for Li+ in ACN (np = 2) indicates that the Li+ cations are less solvated in ACN than the solvents with higher DN, such as in DMSO (np = 4). In other words, the more unshielded Li+ cations in ACN may lead to a stronger attraction to the PS anions. As a result, the chemical equilibrium is expected to shift towards the bound LiPS ion pairs and their respective reaction pathways. This could explain the experimental observation that ACN, despite its intrinsic high dielectric constant (ε = 35.69), behaves like DOL:DME with dominant species S82− and S42−, or more accurately, the Li2S8 and Li2S4 [55].

3. Computational Details

Density functional theory (DFT) calculations of molecules are performed by the Gaussian 09 program [56]. The geometries of polysulfides, lithium polysulfides, solvent molecules and [Li-(solvent)n]+ (n = 1–4) are calculated at the UM11/6 − 311 + G** level [57]. The implicit SMD [58] continuous solvation model is applied to simulate the five selected solvent environments with different dielectric constants (ε), including 1,4-dioxane (DIOX, ε = 2.21), tetrahydrofuran (THF, ε = 7.43), acetone (ε = 20.49), acetonitrile (ACN, ε = 35.69), and dimethyl sulfoxide (DMSO, ε = 46.83).
The one-electron thermodynamic reduction potentials (Ered) are calculated by the following equation [59,60]:
E red   ( Li / Li + ) = G sol F 1.46
where G sol represents the Gibbs free energy of reduction in the solvated environment, F is the Faraday constant and 1.46 V is the absolute reduction electrode potential of Li/Li+. The reaction free energies (∆G) are calculated at 298 K. The average binding energies of explicit solvent molecules with Li+ ( E b ¯ ) are calculated by the following equation:
E b ¯ = E tot [ E Li + + nE solvent ] n
where E tot , E Li + , and E solvent are the computed free energies of the solvated lithium ion-solvent complex, the solvated lithium cation and the solvent molecules, respectively.

4. Conclusions

We perform computational investigations to quantify the influences of the DN and the ε of the solvent system to the relative stability of the polysulfide intermediates in Li-S batteries. The theoretical reduction mechanisms of the PSs under different solvation situations in different solvent systems are shown in Scheme 1. In general, the higher the DN and ε of the solvent, the more favorable the dissociated PS to the bound LiPS, and vice versa. Our theoretical calculations predict different major species when the reduction reactions proceed in the dissociated PS or the bound LiPS state. Our results confirm and rationalize the experimentally observed stability of S62− and S3 in solvents with high DN and high ε, and the observed stability of Li2S8 and Li2S4 in solvents with low DN and low ε. For solvents with a high ε but a low DN, e.g., acetone, the two states shall exist in a chemical equilibrium and the dominant species will depend on the competing forces between solvation and coordination. We hope these results offer a clearer microscopic picture and will inspire the future optimization of the electrolyte systems.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/10/8/911/s1, Figure S1: Optimized geometries of (a) neutral, (b-d) singly-charged, and (e-i) doubly-charged polysulfides in DMSO. Figure S2: Optimized geometries of (a) [Li-(ACN)n]+ and (b) [Li-(DMSO)n]+ clusters with different solvent coordination numbers (n). Table S1: Theoretical values and the reported experimental results of reduction potentials (∆E, in V) of S8, S82-, and S3•- in different solvents.

Author Contributions

G.-Y.D. performed the theoretical calculations; G.-Y.D. and C.-Y.L. wrote the manuscript draft; E.Y.L. supervised the research and revised the manuscript. All authors discussed together. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ministry of Science and Technology (MOST) in Taiwan (MOST 106-2113-M-003-010-MY3).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yin, Y.-X.; Xin, S.; Guo, Y.-G.; Wan, L.-J. Lithium–sulfur batteries: Electrochemistry, materials, and prospects. Angew. Chem. Int. Ed. 2013, 52, 13186–13200. [Google Scholar] [CrossRef] [PubMed]
  2. Ould Ely, T.; Kamzabek, D.; Chakraborty, D.; Doherty, M.F. Lithium-sulfur batteries: State of the art and future directions. ACS Appl. Energy Mater. 2018, 1, 1783–1814. [Google Scholar] [CrossRef]
  3. Tarascon, J.-M.; Armand, M. Issues and challenges facing rechargeable lithium batteries. Nature 2001, 414, 359–367. [Google Scholar] [CrossRef] [PubMed]
  4. Kang, W.; Deng, N.; Ju, J.; Li, Q.; Wu, D.; Ma, X.; Li, L.; Naebe, M.; Cheng, B. A review of recent developments in rechargeable lithium–sulfur batteries. Nanoscale 2016, 8, 16541–16588. [Google Scholar] [CrossRef] [PubMed]
  5. Wild, M.; O’Neill, L.; Zhang, T.; Purkayastha, R.; Minton, G.; Marinescu, M.; Offer, G.J. Lithium sulfur batteries, a mechanistic review. Energy Environ. Sci. 2015, 8, 3477–3494. [Google Scholar] [CrossRef]
  6. Eroglu, D.; Zavadil, K.R.; Gallagher, K.G. Critical link between materials chemistry and cell-level design for high energy density and low cost lithium-sulfur transportation battery. J. Electrochem. Soc. 2015, 162, A982. [Google Scholar] [CrossRef]
  7. Hagen, M.; Hanselmann, D.; Ahlbrecht, K.; Maca, R.; Gerber, D.; Tubke, J. Lithium-sulfur cells: The gap between the state-of-the-art and the requirements for high energy battery cells. Adv. Energy Mater. 2015, 5, 1401986. [Google Scholar] [CrossRef]
  8. Salama, M.; Rosy Attias, R.; Yemini, R.; Gofer, Y.; Aurbach, D.; Noked, M. Metal–sulfur batteries: Overview and research methods. ACS Energy Lett. 2019, 4, 436–446. [Google Scholar] [CrossRef]
  9. Fang, R.; Zhao, S.; Sun, Z.; Wang, D.-W.; Cheng, H.-M.; Li, F. More reliable lithium-sulfur batteries: Status, solutions and prospects. Adv. Mater. 2017, 29, 1606823. [Google Scholar] [CrossRef]
  10. Wang, H.; Sa, N.; He, M.; Liang, X.; Nazar, L.F.; Balasubramanian, M.; Gallagher, K.G.; Key, B. In situ NMR observation of the temporal speciation of lithium sulfur batteries during electrochemical cycling. J. Phys. Chem. C 2017, 121, 6011–6017. [Google Scholar] [CrossRef]
  11. Busche, M.R.; Adelhelm, P.; Sommer, H.; Schneider, H.; Leitner, K.; Janek, J. Systematical electrochemical study on the parasitic shuttle-effect in lithium-sulfur-cells at different temperatures and different rates. J. Power Sources 2014, 259, 289–299. [Google Scholar] [CrossRef]
  12. Xu, N.; Qian, T.; Liu, X.; Liu, J.; Chen, Y.; Yan, C. Greatly suppressed shuttle effect for improved lithium sulfur battery performance through short chain intermediates. Nano Lett. 2017, 17, 538–543. [Google Scholar] [CrossRef]
  13. Lin, D.; Liu, Y.; Cui, Y. Reviving the lithium metal anode for high-energy batteries. Nat. Nanotechnol. 2017, 12, 194. [Google Scholar] [PubMed]
  14. Bruce, P.G.; Freunberger, S.A.; Hardwick, L.J.; Tarascon, J.-M. Li–O2 and Li–S batteries with high energy storage. Nat. Mater. 2012, 11, 19. [Google Scholar] [CrossRef] [PubMed]
  15. Carbone, L.; Coneglian, T.; Gobet, M.; Munoz, S.; Devany, M.; Greenbaum, S.; Hassoun, J. A simple approach for making a viable, safe, and high-performances lithium-sulfur battery. J. Power Sources 2018, 377, 26–35. [Google Scholar] [CrossRef]
  16. Angulakshmi, N.; Stephan, A.M. Efficient electrolytes for lithium–sulfur batteries. Front. Energy Res. 2015, 3, 17. [Google Scholar] [CrossRef] [Green Version]
  17. Younesi, R.; Veith, G.M.; Johansson, P.; Edstrom, K.; Vegge, T. Lithium salts for advanced lithium batteries: Li–metal, Li–O 2, and Li–S. Energy Environ. Sci. 2015, 8, 1905–1922. [Google Scholar] [CrossRef] [Green Version]
  18. Xu, J.; Zhang, W.; Fan, H.; Cheng, F.; Su, D.; Wang, G. Promoting lithium polysulfide/sulfide redox kinetics by the catalyzing of zinc sulfide for high performance lithium-sulfur battery. Nano Energy 2018, 51, 73–82. [Google Scholar] [CrossRef]
  19. Li, X.; Cao, Y.; Qi, W.; Saraf, L.V.; Xiao, J.; Nie, Z.; Mietek, J.; Zhang, J.-G.; Stchwenzer, B.; Liu, J. Optimization of mesoporous carbon structures for lithium–sulfur battery applications. J. Mater. Chem. 2011, 21, 16603–16610. [Google Scholar] [CrossRef]
  20. Ghosh, A.; Garapati, M.S.; Saroja, A.P.V.K.; Sundara, R. Polar Bilayer Cathode for Advanced Lithium–Sulfur Battery: Synergy Between Polysulfide Conversion and Confinement. J. Phys. Chem. C 2019, 123, 10777–10787. [Google Scholar] [CrossRef]
  21. Fang, X.; Peng, H. A revolution in electrodes: Recent progress in rechargeable lithium–sulfur batteries. Small 2015, 11, 1488–1511. [Google Scholar] [CrossRef] [PubMed]
  22. Paolella, A.; Demers, H.; Chevallier, P.; Gagnon, C.; Girard, G.; Delaporte, N.; Zhu, W.; Vigh, A.; Guerfi, A.; Zaghib, K. A platinum nanolayer on lithium metal as an interfacial barrier to shuttle effect in Li-S batteries. J. Power Sources 2019, 427, 201–206. [Google Scholar] [CrossRef]
  23. Paolella, A.; Laul, D.; Timoshevskii, V.; Zhu, W.; Marras, S.; Bertoni, G.; Wahba, A.S.; Girard, G.; Gagnon, C.; Rodrigue, L.; et al. The role of metal disulfide interlayer in Li–S batteries. J. Phys. Chem. C 2018, 122, 1014–1023. [Google Scholar] [CrossRef]
  24. Liu, J.; HYuan, L.; Yuan, K.; Li, Z.; Hao, Z.; Xiang, J.; Huang, Y. SnO2 as a high-efficiency polysulfide trap in lithium–sulfur batteries. Nanoscale 2016, 8, 13638–13645. [Google Scholar] [CrossRef] [PubMed]
  25. Wu, F.; Ye, Y.; Chen, R.; Qian, J.; Zhao, T.; Li, L.; Li, W. Systematic effect for an ultralong cycle lithium–sulfur battery. Nano Lett. 2015, 15, 7431–7439. [Google Scholar] [CrossRef] [PubMed]
  26. Lochala, J.; Liu, D.; Wu, B.; Robinson, C.; Xiao, J. Research progress toward the practical applications of lithium–sulfur batteries. ACS Appl. Mater. Inter. 2017, 9, 24407–24421. [Google Scholar] [CrossRef]
  27. Chen, X.; Hou, T.; Persson, K.A.; Zhang, Q. Combining theory and experiment in lithium–sulfur batteries: Current progress and future perspectives. Mater. Today 2019, 22, 142–158. [Google Scholar] [CrossRef]
  28. Park, C.; Ronneburg, A.; Risse, S.; Ballauff, M.; Kanduc, M.; Dzubiella, J. Structural and Transport Properties of Li/S Battery Electrolytes: Role of the Polysulfide Species. J. Phys. Chem. C 2019, 123, 10167–10177. [Google Scholar] [CrossRef]
  29. Kamphaus, E.P.; Balbuena, P.B. Long-chain polysulfide retention at the cathode of Li–S batteries. J. Phys. Chem. C 2016, 120, 4296–4305. [Google Scholar] [CrossRef]
  30. Park, C.; Kanduc, M.; Chudoba, R.; Ronneburg, A.; Risse, S.; Ballauff, M.; Dzubiella, J. Molecular simulations of electrolyte structure and dynamics in lithium–sulfur battery solvents. J. Power Sources 2018, 373, 70–78. [Google Scholar] [CrossRef] [Green Version]
  31. Camacho-Forero, L.E.; Smith, T.W.; Bertolini, S.; Balbuena, P.B. Reactivity at the lithium–metal anode surface of lithium–sulfur batteries. J. Phys. Chem. C 2015, 119, 26828–26839. [Google Scholar] [CrossRef]
  32. Liu, Z.; Bertolini, S.; Balbuena, P.B.; Mukherjee, P.P. Li2S film formation on lithium anode surface of Li–S batteries. ACS Appl. Mater. Interfaces 2016, 8, 4700–4708. [Google Scholar] [CrossRef] [PubMed]
  33. Pascal, T.A.; Wujcik, K.H.; Velasco-Velez, J.; Wu, C.; Teran, A.A.; Kapilashrami, M.; Cabana, J.; Guo, J.; Salmeron, M.; Balsara, N.; et al. X-ray absorption spectra of dissolved polysulfides in lithium–sulfur batteries from first-principles. J. Phys. Chem. Lett. 2014, 5, 1547–1551. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Kawase, A.; Shirai, S.; Yamoto, Y.; Arakawa, R.; Takata, T. Electrochemical reactions of lithium–sulfur batteries: An analytical study using the organic conversion technique. Phys. Chem. Chem. Phys. 2014, 16, 9344–9350. [Google Scholar] [CrossRef] [PubMed]
  35. Lu, Y.-C.; He, Q.; Gasteiger, H.A. Probing the lithium–sulfur redox reactions: A rotating-ring disk electrode study. J. Phys. Chem. C 2014, 118, 5733–5741. [Google Scholar] [CrossRef]
  36. Assary, R.S.; Curtiss, L.A.; Moore, J.S. Toward a molecular understanding of energetics in Li–S batteries using nonaqueous electrolytes: A high-level quantum chemical study. J. Phys. Chem. C 2014, 118, 11545–11558. [Google Scholar] [CrossRef]
  37. See, K.A.; Wu, H.-L.; Lau, K.C.; Shin, M.; Cheng, L.; Balasubramanian, M.; Gallagher, K.G.; Curtiss, L.A.; Gewirth, A.A. Effect of hydrofluoroether cosolvent addition on Li solvation in acetonitrile-based solvate electrolytes and its influence on S reduction in a Li–S battery. ACS Appl. Mater. Interfaces 2016, 8, 34360–34371. [Google Scholar] [CrossRef]
  38. Zou, Q.; Lu, Y.-C. Solvent-dictated lithium sulfur redox reactions: An operando UV–vis spectroscopic study. J Phys. Chem. Lett. 2016, 7, 1518–1525. [Google Scholar] [CrossRef]
  39. Pascal, T.A.; Wujcik, K.H.; Wang, D.R.; Balsara, N.P.; Prendergast, D. Thermodynamic origins of the solvent-dependent stability of lithium polysulfides from first principles. Phys. Chem. Chem. Phys. 2017, 19, 1441–1448. [Google Scholar] [CrossRef] [Green Version]
  40. Wu, H.-L.; Huff, L.A.; Gewirth, A.A. In situ Raman spectroscopy of sulfur speciation in lithium–sulfur batteries. ACS Appl. Mater. Interfaces 2015, 7, 1709–1719. [Google Scholar] [CrossRef]
  41. Vijayakumar, M.; Govind, N.; Walter, E.; Burton, S.D.; Shukla, A.; Devaraj, A.; Xiao, J.; Liu, J.; Wang, C.; Karim, A.; et al. Molecular structure and stability of dissolved lithium polysulfide species. Phys. Chem. Chem. Phys. 2014, 16, 10923–10932. [Google Scholar] [CrossRef] [PubMed]
  42. Cheng, L.; Curtiss, L.A.; Zavadil, K.R.; Gewirth, A.A.; Shao, Y.; Gallagher, K.G. Sparingly solvating electrolytes for high energy density lithium–sulfur batteries. ACS Energy Lett. 2016, 1, 503–509. [Google Scholar] [CrossRef]
  43. Wujcik, K.H.; Wang, D.R.; Raghunathan, A.; Drake, M.; Pascal, T.A.; Prendergast, D.; Balsara, N.P. Lithium polysulfide radical anions in ether-based solvents. J. Phys. Chem. C 2016, 120, 18403–18410. [Google Scholar] [CrossRef] [Green Version]
  44. Steudel, R.; Chivers, T. The role of polysulfide dianions and radical anions in the chemical, physical and biological sciences, including sulfur-based batteries. Chem. Soc. Rev. 2019, 48, 3279–3319. [Google Scholar] [CrossRef] [PubMed]
  45. Cuisinier, M.; Hart, C.; Balasubramanian, M.; Garsuch, A.; Nazar, L.F. Radical or not radical: Revisiting lithium–sulfur electrochemistry in nonaqueous electrolytes. Adv. Energy Mater. 2015, 5, 1401801. [Google Scholar] [CrossRef]
  46. Zou, Q.; Liang, Z.; Du, G.-Y.; Liu, C.-Y.; Li, E.Y.; Lu, Y.-C. Cation-directed selective polysulfide stabilization in alkali metal–sulfur batteries. J. Am. Chem. Soc. 2018, 140, 10740–10748. [Google Scholar] [CrossRef]
  47. Li, G.; Li, Z.; Zhang, B.; Lin, Z. Developments of electrolyte systems for lithium–sulfur batteries: A review. Front. Energy Res. 2015, 3, 5. [Google Scholar] [CrossRef] [Green Version]
  48. Wujcik, K.H.; Pascal, T.A.; Pemmaraju, C.D.; Devaux, D.; Stolte, W.C.; Balsara, N.P.; Prendergast, D. Characterization of Polysulfide Radicals Present in an Ether-Based Electrolyte of a Lithium–Sulfur Battery During Initial Discharge Using In Situ X-Ray Absorption Spectroscopy Experiments and First-Principles Calculations. Adv. Energy Mater 2015, 5, 1500285. [Google Scholar] [CrossRef] [Green Version]
  49. Wang, B.; Alhassan, S.M.; Pantelides, S.T. Formation of large polysulfide complexes during the lithium-sulfur battery discharge. Phys. Rev. Appl. 2014, 2, 034004. [Google Scholar] [CrossRef]
  50. Xin, S.; Gu, L.; Zhao, N.-H.; Yin, Y.-X.; Zhou, L.-J.; Guo, Y.-G.; Wan, L.-J. Smaller sulfur molecules promise better lithium–sulfur batteries. J. Am. Chem. Soc. 2012, 134, 18510–18513. [Google Scholar] [CrossRef]
  51. Steudel, R.; Steudel, Y. Polysulfide Chemistry in Sodium-Sulfur Batteries and Related Systems–A Computational Study by G3X (MP2) and PCM Calculations. Chem.: Eur. J. 2013, 19, 3162–3176. [Google Scholar] [CrossRef] [PubMed]
  52. Patel, M.U.M.; Dominko, R. Application of in operando UV/vis spectroscopy in lithium–sulfur batteries. ChemSusChem 2014, 7, 2167–2175. [Google Scholar] [CrossRef] [PubMed]
  53. Patel, M.U.M.; Demir-Cakan, R.; Morcrette, M.; Tarascon, J.-M.; Gaberscek, M.; Dominko, R. Li-S Battery Analyzed by UV/Vis in Operando Mode. ChemSusChem 2013, 6, 1177–1181. [Google Scholar] [CrossRef] [PubMed]
  54. Fawcett, W.R. Acidity and basicity scales for polar solvents. J. Phys. Chem. 1993, 97, 9540–9546. [Google Scholar] [CrossRef]
  55. He, Q.; Gorlin, Y.; Patel, M.U.; Gasteiger, H.A.; Lu, Y.-C. Unraveling the correlation between solvent properties and sulfur redox behavior in lithium-sulfur batteries. J. Electrochem. Soc. 2018, 165, A4027. [Google Scholar] [CrossRef]
  56. 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 09; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  57. Peverati, R.; Truhlar, D.G. Improving the accuracy of hybrid meta-GGA density functionals by range separation. J. Phys. Chem. Lett. 2011, 2, 2810–2817. [Google Scholar] [CrossRef]
  58. 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]
  59. Sriana, T.; Leggesse, E.G.; Jiang, J.C. Novel benzimidazole salts for lithium ion battery electrolytes: Effects of substituents. Phys. Chem. Chem. Phys. 2015, 17, 16462–16468. [Google Scholar] [CrossRef]
  60. 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]
Figure 1. Optimized geometries of different LiPSs in DMSO. The average bond lengths of S-S (d, in Å) and the bond angles (θ, in °) are labeled in red and purple, respectively.
Figure 1. Optimized geometries of different LiPSs in DMSO. The average bond lengths of S-S (d, in Å) and the bond angles (θ, in °) are labeled in red and purple, respectively.
Catalysts 10 00911 g001
Scheme 1. Schematic presentation for the solvent interaction with the PS or LiPS species and the corresponding reduction reaction mechanisms in different solvent systems.
Scheme 1. Schematic presentation for the solvent interaction with the PS or LiPS species and the corresponding reduction reaction mechanisms in different solvent systems.
Catalysts 10 00911 sch001
Table 1. The reaction free energies (∆G, eV) of polysulfides in different solvents. The dielectric constants (ε) of different solvents are presented in parentheses. The solvents that tend to facilitate the reactions of the dissociated PS’s are shaded in grey, and the reaction energies for the more probable reactions are in bold.
Table 1. The reaction free energies (∆G, eV) of polysulfides in different solvents. The dielectric constants (ε) of different solvents are presented in parentheses. The solvents that tend to facilitate the reactions of the dissociated PS’s are shaded in grey, and the reaction energies for the more probable reactions are in bold.
#ReactionsDIOX
(2.21)
THF
(7.43)
Acetone
(20.49)
ACN
(35.69)
DMSO
(46.83)
1S82−→S62− + S20.470.380.310.320.32
2S82-→S5 + S3−0.87−0.080.160.230.26
3S82−→S42− + S41.631.251.121.111.11
4S82−→2S4−0.760.040.230.290.31
5S62−→2S3−1.30−0.45−0.19−0.12−0.10
6S62−→S42− + S20.980.700.630.620.62
72S42−→2S3 + S22−−0.270.390.600.660.69
82S42−→S62− + S22−1.030.840.800.780.78
Table 2. The reaction free energies (∆G, eV) of lithium polysulfides in different solvents. The dielectric constants (ε) of different solvents are presented in parentheses. The solvents that tend to facilitate the reactions of the bound LiPS’s are shaded in grey, and the reaction energies for the more probable reactions are in bold.
Table 2. The reaction free energies (∆G, eV) of lithium polysulfides in different solvents. The dielectric constants (ε) of different solvents are presented in parentheses. The solvents that tend to facilitate the reactions of the bound LiPS’s are shaded in grey, and the reaction energies for the more probable reactions are in bold.
#ReactionsDIOX
(2.21)
THF
(7.43)
Acetone
(20.49)
ACN
(35.69)
DMSO
(46.83)
9Li2S8→Li2S6 + S20.200.190.170.180.19
10Li2S8→LiS5 + LiS31.130.950.850.800.91
11Li2S8→Li2S4 + S40.850.810.790.790.81
12Li2S8→2LiS41.150.970.890.870.87
13Li2S6→2LiS30.880.670.570.480.70
14Li2S6→Li2S4 + S20.470.450.440.450.46
152Li2S4→2LiS3 + Li2S22.111.711.371.211.12
162Li2S4→Li2S6 + Li2S20.970.820.700.650.64
Table 3. The predicted coordination number (np) as well as the average binding energies ( E b ¯ ) with different number of binding solvent molecules (n) on Li+ cation. The donor number (DN) and the dielectric constants (ε) of different solvents are also listed.
Table 3. The predicted coordination number (np) as well as the average binding energies ( E b ¯ ) with different number of binding solvent molecules (n) on Li+ cation. The donor number (DN) and the dielectric constants (ε) of different solvents are also listed.
SolventDNεnp E b ¯
n = 1n = 2n = 3n = 4
ACN14.135.692−0.40−0.35−0.29−0.22
Acetone17.020.493−0.51−0.45−0.31−0.25
DMSO29.846.834−0.77−0.67−0.52−0.40

Share and Cite

MDPI and ACS Style

Du, G.-Y.; Liu, C.-Y.; Li, E.Y. A DFT Investigation on the Origins of Solvent-Dependent Polysulfide Reduction Mechanism in Rechargeable Li-S Batteries. Catalysts 2020, 10, 911. https://doi.org/10.3390/catal10080911

AMA Style

Du G-Y, Liu C-Y, Li EY. A DFT Investigation on the Origins of Solvent-Dependent Polysulfide Reduction Mechanism in Rechargeable Li-S Batteries. Catalysts. 2020; 10(8):911. https://doi.org/10.3390/catal10080911

Chicago/Turabian Style

Du, Guan-Ying, Chi-You Liu, and Elise Y. Li. 2020. "A DFT Investigation on the Origins of Solvent-Dependent Polysulfide Reduction Mechanism in Rechargeable Li-S Batteries" Catalysts 10, no. 8: 911. https://doi.org/10.3390/catal10080911

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop