Skip to main content
ACS AuthorChoice logoLink to ACS AuthorChoice
. 2021 Aug 27;12(35):8460–8464. doi: 10.1021/acs.jpclett.1c02474

Importance of the Ion-Pair Lifetime in Polymer Electrolytes

Harish Gudla 1, Yunqi Shao 1, Supho Phunnarungsi 1, Daniel Brandell 1, Chao Zhang 1,*
PMCID: PMC8436209  PMID: 34449227

Abstract

graphic file with name jz1c02474_0005.jpg

Ion pairing is commonly considered as a culprit for the reduced ionic conductivity in polymer electrolyte systems. However, this simple thermodynamic picture should not be taken literally, as ion pairing is a dynamical phenomenon. Here we construct model poly(ethylene oxide)–bis(trifluoromethane)sulfonimide lithium salt systems with different degrees of ion pairing by tuning the solvent polarity and examine the relation between the cation–anion distinct conductivity σ+–d and the lifetime of ion pairs τ+– using molecular dynamics simulations. It is found that there exist two distinct regimes where σ+– scales with 1/τ+– and τ+–, respectively, and the latter is a signature of longer-lived ion pairs that contribute negatively to the total ionic conductivity. This suggests that ion pairs are kinetically different depending on the solvent polarity, which renders the ion-pair lifetime highly important when discussing its effect on ion transport in polymer electrolyte systems.


Ion pairing in electrolyte solutions16 results from a delicate balance between ion–solvent and ion–ion interactions. One common approach to define an ion pair is to use Bjerrum’s criterion,7 in which the distance r+– is smaller than the effective range of −q+q/2ε kbT (half of the Bjerrum length) with ε as the dielectric constant of the solution, q+ and q being the ionic charges, the Boltzmann constant kb, and the temperature T. Bjerrum’s criterion suggests that the solvent polarity plays a critical role in the formation of ion pairs, rendering a distinction between contact ion pairs (CIPs) and solvent-separated ion pairs (SSIPs).1 In addition, it implies that the formation of pairs of equal ionic species is unlikely to occur due to the electrostatic repulsion but that the possibility of forming triplets or larger aggregates, for example, an anion–cation–anion cluster, cannot be excluded.8,9

The idea that ion pairing affects the ionic conductivity was introduced early on by Arrhenius, who ascribed the decrease of the equivalent conductivity at a higher concentration to the formation of charge-neutral ion pairs.10 This idea has been put forward using the molar conductivity ratio ΛEISNMR measured by electrochemical impedance spectroscopy (EIS) and pulse-field gradient NMR to quantify the ionicity (the degree of dissociativity), particularly for ionic liquids11,12 and polymer electrolytes.13 Nevertheless, it has been realized that deviations of the ionic conductivity from the Nernst–Einstein relation cannot solely be attributed to the formation of ion pairs,1416 where other factors such as the hydrodynamic interactions manifested via viscosity can play an important role.17

To describe the effect of ion pairing on the ionic conductivity, one needs an observable that can be accessed both theoretically and experimentally. The key quantity used here is the cation–anion distinct conductivity σ+–d from liquid-state theory18,19

graphic file with name jz1c02474_m001.jpg 1

where Ω is the volume of the system, and Δr(t) is the displacement vector of each ion at time t. Note that σ+–d is experimentally measurable20,21 and directly related to the Onsager transport coefficient Ω+–.19,22

Somewhat unexpectedly, σ+–d is often found positive (instead of negative as in Arrhenius’ picture) in different types of electrolyte systems, spanning categories from aqueous electrolyte solutions to polymer ionic liquids.15,17,2327 This suggests that the existence of ion pairs, as evinced by a number of spectroscopic experiments,2830 does not necessarily imply a negative contribution to the measured ionic conductivity but can instead contribute to an increase in the transport of ions. Therefore, understanding the ion pairing effect on polymer electrolytes is crucial, as their application in energy storage systems is largely limited by a low ionic conductivity.3133

The crucial point to this conundrum lies in the fact that Bjerrum’s convention is a thermodynamic criterion, while the ionic conductivity is a dynamical property. Therefore, the lifetime of charge-neutral ion pairs needs to be considered explicitly when discussing the contribution of ion pairing to the ionic conductivity, in addition to the distance criterion due to the thermodynamic stability. In other words, an ion pair should be “long-lived enough to be a recognizable kinetic entity”.34

Theoretically, the lifetime of ion pairs τ+– can be extracted from the normalized time correlation function of the cation–anion pairs in molecular dynamics (MD) simulations35

graphic file with name jz1c02474_m002.jpg 2

where f(rij;s) is a window function to detect whether a cation–anion pair lies within the cutoff rc for a given period s.

The first approach is to use the product of the Heaviside functions θ(x) defined by a time series of pairwise distances rij between a cation–anion pair, as follows.36

graphic file with name jz1c02474_m003.jpg 3

However, the persistence time (PT) from this procedure clearly neglects recrossing events, for example, reactions passing over the transition state but returning to the reactant afterward, which has been discussed extensively for hydrogen-bond dynamics.35,37 Here, we used the stable states picture (SSP) of chemical reactions proposed by Laage and Hynes, which remedies this problem.38 Then f(rij; s) in SSP is given as

graphic file with name jz1c02474_m004.jpg 4

where rc,prod is the product SSP boundary, which corresponds to the cation–anion distance at the half height of the second peak in the radial distribution function (RDF). Then, rc in eq 2 should be replaced by the reactant SSP boundary rc,reac, which is at the first maximum of the cation–anion RDF.

To investigate the relation between σ+–d and τ+– in polymer electrolyte systems, we constructed simulation boxes consisting of 200 poly(ethylene oxide) (PEO) polymer chains each with 25 ethylene oxide (EO) repeating units and 400 bis(trifluoromethane)sulfonimide lithium salt (LiTFSI) ([Li+]/[EO] concentration = 0.08). As indicated by Bjerrum’s criterion, the solvent polarity strongly modulates the ion pairing. This motivated us to apply the charge scaling method39 to PEO molecules to change the degree of ion pairing. General AMBER force field (GAFF)40 parameters were used for describing bonding and nonbonding interactions in PEO and LiTFSI, and all MD simulations were performed using GROMACS 2018.1.41 All systems were properly equilibrated to make sure that the simulation length is larger than the Rouse time of the polymer. Details for the system setup and MD simulations can be found in the Supporting Information Section A.1.

Before discussing our main result, it is necessary to check how structural and transport properties change when we tweak the handle of the solvent polarity. Here, the solvent polarity is described by the dielectric constant of the system εP, which was computed for each polymer electrolyte system (see Section A.2 in the Supporting Information for details).

The RDFs of Li–N(TFSI) are plotted in Figure 1a, where peaks in the Li–N(TFSI) RDF increase significantly when εP becomes smaller. This is a sign of formation of ion pairs, which is also evinced in Figure 2. Accordingly, there is an optimal value in the total Green–Kubo conductivity σG–K when the solvent polarity is modulated as seen in Figure 1b. Both of these results support our previous observations of the effect of solvent polarity on the Li+ transportation in PEO-LiTFSI systems42 and agree with other recent studies of polymer electrolyte systems.43,44

Figure 1.

Figure 1

(a) The Li–N(TFSI) RDF at different solvent polarity strengths (as quantified by the dielectric constant εP). (b) The total conductivity σG-K computed from the Green–Kubo relation as a function of εP. (c) The lifetime of ion pairs τ+– computed from the SSP method as a function of εP, where rc,reac = 2.1 Å and rc,prod = 3.8–5.5 Å. (d) The cation–anion distinct conductivity σ+–d (and its decomposition into pairing and non-pairing contributions) as a function of εP.

Figure 2.

Figure 2

Modulation of ion pairing in PEO-LiTFS systems by solvent polarity εP. Sky blue - PEO chains, Purple - Li ions, Orange - TFSI ions.

Figure 1c,d, however, demonstrates novel phenomena. The lifetime of ion pairs increases when εP is either high or low, and it reaches a minimum at the intermediate value of εP (see Section A.3 in the Supporting Information for further details of calculations of the lifetime of ion pairs and a comparison of outcomes from eq 3 and eq 4). Inspecting Figure 1b,c, one may attempt to relate the opposite trend shown in the total ionic conductivity σG–K to that of τ+–. However, the lifetime increases much more rapidly at a lower dielectric constant regime (εP < 2.3) compared to that at a higher dielectric constant regime (εP > 3). This suggests there are different types of ion pairs in polymer electrolyte systems under investigation here. Looking at the cation–anion distinct conductivity σ+–d, one can clearly see that it goes from positive to negative when εP becomes smaller. (Note that σ+– > 0 corresponds to anticorrelated cation–anion movements for the sign convention used in this work.) In particular, the rapid decrement in σ+–d at lower εP seems in accord with the rapid increment in τ+–. These observations also point to the direction that these two key properties of ion pairs, namely, σ+– and τ+–, must be closely related.

This leads to our main result shown in Figure 3. What we find is that there exist two distinct regimes: σ+–d scales with 1/τ+– (for higher values of εP), and σ+– scales with τ+– (for lower values of εP). Moreover, the transition between these two regimes shows a combined feature. Therefore, the general scaling relation we propose for polymer electrolyte systems is

graphic file with name jz1c02474_m005.jpg 5

where both A and B are system-dependent coefficients. Therefore, what matters to discussions of the ion-pairing effect on transport properties in polymer electrolytes is not whether ion pairs are present or not in the system but how long they live. By establishing the scaling relation for ion pairs from MD simulations, one could predict the lifetime of ion pairs using the measured value of σ+–d in experiments.27

Figure 3.

Figure 3

Scaling relation between the cation–anion distinct conductivity σ+–d and the lifetime of ion pairs τ+– computed from the SSP method for PEO-LiTFSI polymer electrolyte systems with different solvent polarity strengths.

Then, the immediate question that appears is, Why do shorter-lived ion pairs scale with 1/τ+–, while longer-lived counterparts scale with τ+–? The first scaling relation seems rather general, as already observed in ionic liquids, organic electrolytes, polymer ionic liquids, and salt-doped homopolymers.23,26,4547 This is reminiscent of the Walden rule or the Stokes–Einstein relation. The second scaling relation we find in this work is a consequence of that τ+– computed from the SSP method is equal to the inverse of the reactive flux rate constant 1/kRF38 for the ion-pair dissociation, and 1/kRF is proportional to the concentration of ion pairs from the law of mass action. Since σ+–d is also proportional to the number of ion pairs, this leads to the observation that σ+– scales linearly with τ+–.

It is worth mentioning that σ+–d includes both contributions from the longer-lived ion pairs and the remainder. This suggests that one could further separate these two contributions for longer-lived pairs

graphic file with name jz1c02474_m006.jpg 6

where fSSP is the same function given by eq 4. Then, the contribution from the remainder is simply σ+–d, nonpairing = σ+– – σ+–d, pairing. Here, the parameter s is chosen to be 2 ns, as decided by the convergence of the conductivity calculation (see Section A.4 in the Supporting Information).

The result of this decomposition is shown in Figure 1d. The σ+–d, pairing remains zero until a lower value of εP. This agrees with the appearance of longer-lived ion pairs as seen in Figure 1c. More interestingly, in the presence of longer-lived ion pairs, the σ+– is negative, but the σ+–d, nonpairing is positive instead. To understand why, we made a toy model of a NaCl solution where all Na–Cl are paired up with holonomic constraints. Details for the system setup and MD simulations can be found in Section B of the Supporting Information.

Mean square charge displacements (MSCD, i.e., quantities inside the square bracket in eq 1 and eq 6) of this toy model are shown in Figure 4, for the total ionic conductivity σG-K, self-conductivities (σ+ + σ), and the sum of cation–cation and anion–anion distinct conductivities (σ++d + σ––) as well as σ+–d, pairing and σ+–. Since all Na–Cl ion pairs are permanent by construction, the total ionic conductivity as the sum of all these individual contributions mentioned above must be zero (i.e., the slope of the MSCD “total” is zero), as evinced in Figure 4. Moreover, self-conductivities (σ+ + σ) should be exactly the negative of the direct part of the cation–anion distinct conductivity σ+–d, pairing, as seen also in Figure 4. On the basis of these considerations, we know that the sum of σ++, σ––d, and σ+– is zero as well. As shown in Figure 4, σ++d and σ–– are negative, while σ+–d, nonpairing is positive in the toy model with permanent ion pairs. This provides a rationale to the opposite signs of σ+– and σ+–d, nonpairing as seen in Figure 1d of PEO-LiTFSI systems. Nevertheless, one should be aware that the situation with ion aggregates will be different, as σ++ and σ––d could be positive instead.

Figure 4.

Figure 4

MSCD of a 5 mol NaCl solution with permanent ion pairs (in the same order as those that appeared in the key box): for the total ionic conductivity σG-K, the self-conductivity (σ+ + σ), the sum of cation–cation and anion–anion distinct conductivities (σ++d + σ––), the ion pairing part of the cation–anion distinct conductivity σ+–d, pairing, and the remainder part σ+–. N is the number of ion pairs in the system, and μ is the dipole moment of each ion pair.

To sum up, following Bjerrum’s criterion, we have constructed PEO-LiTFSI systems in our MD simulations with different degrees of ion pairing by modulating the solvent polarity. What we found is that there exist two distinct regimes where the cation–anion distinct conductivity σ+–d scales with 1/τ+– and with τ+–, respectively. The linear scaling of σ+– with respect to the lifetime of the ion pairs τ+– is a signature of longer-lived ion pairs that reduce the total ionic conductivity. By establishing this scaling relation, one could infer the lifetime of ion pairs from the experimentally measured cation–anion distinct conductivity. This further suggests that what matters to discussions of the ion-pairing effect on transport properties in polymer electrolytes is not the presence of ion pairs but the corresponding lifetime.

In the scaling relation we found in our MD simulations (eq 5), the coefficient A is positive, which suggests anticorrelated movements of cation–anions and shorter-lived ion pairs. This hints that ion aggregates analyzed in previous MD studies of the PEO-LiTFSI system48 would not populate in this scenario, in line with conclusions drawn from other experimental works for polymer electrolytes.49,50 Nevertheless, it is worth noting that early experiments on aqueous ionic solutions show that σ+–d can flip the sign from negative (correlated) to positive (anticorrelated) when the salt concentration is increased,20 which is intriguing. This clearly indicates that the cation–anion distinct conductivity σ+– is a sensitive probe to the convoluted ion–ion correlations, which calls for further investigations from both experiments and simulations to understand its nature and its relationship with other static and dynamical properties in electrolyte systems.

Acknowledgments

This work has been supported by the European Research Council (ERC), Grant No. 771777 “FUN POLYSTORE”, and the Swedish Research Council (VR), Grant No. 2019-05012. The authors are thankful for the funding from the Swedish National Strategic e-Science program eSSENCE and STandUP for Energy. The simulations were performed on the resources provided by the Swedish National Infrastructure for Computing (SNIC) at NSC, PDC, and UPPMAX.

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.jpclett.1c02474.

  • Descriptions of the setup and MD simulations of PEO-LiTFSI systems and NaCl solution with permanent ion pairs (PDF)

The authors declare no competing financial interest.

Supplementary Material

jz1c02474_si_001.pdf (574.9KB, pdf)

References

  1. Marcus Y.; Hefter G. Ion pairing. Chem. Rev. 2006, 106, 4585–4621. 10.1021/cr040087x. [DOI] [PubMed] [Google Scholar]
  2. Macchioni A. Ion pairing in transition-metal organometallic chemistry. Chem. Rev. 2005, 105, 2039–2074. 10.1021/cr0300439. [DOI] [PubMed] [Google Scholar]
  3. Rehm T. H.; Schmuck C. Ion-pair induced self-assembly in aqueous solvents. Chem. Soc. Rev. 2010, 39, 3597–3611. 10.1039/b926223g. [DOI] [PubMed] [Google Scholar]
  4. Brak K.; Jacobsen E. N. Asymmetric ion-pairing catalysis. Angew. Chem., Int. Ed. 2013, 52, 534–561. 10.1002/anie.201205449. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Rajput N. N.; Qu X.; Sa N.; Burrell A. K.; Persson K. A. The coupling between stability and ion pair formation in magnesium electrolytes from first-principles quantum mechanics and classical molecular dynamics. J. Am. Chem. Soc. 2015, 137, 1–10. 10.1021/jacs.5b01004. [DOI] [PubMed] [Google Scholar]
  6. Huang Y.; Liang Z.; Wang H. A dual-ion battery has two sides: the effect of ion-pairs. Chem. Commun. 2020, 56, 10070–10073. 10.1039/D0CC03951A. [DOI] [PubMed] [Google Scholar]
  7. Bjerrum N. Untersuchungen über Ionenassoziation. I. Kgl. Dankse Vid. Selskab. Math. -Fys. Medd. 1926, 7, 1–48. [Google Scholar]
  8. Evans J.; Vincent C. A.; Bruce P. G. Electrochemical measurement of transference numbers in polymer electrolytes. Polymer 1987, 28, 2324–2328. 10.1016/0032-3861(87)90394-6. [DOI] [Google Scholar]
  9. Villaluenga I.; Pesko D. M.; Timachova K.; Feng Z.; Newman J.; Srinivasan V.; Balsara N. P. Negative Stefan-Maxwell diffusion coefficients and complete electrochemical transport characterization of homopolymer and block copolymer electrolytes. J. Electrochem. Soc. 2018, 165, A2766–A2773. 10.1149/2.0641811jes. [DOI] [Google Scholar]
  10. Fawcett W. R.Liquids, solutions, and interfaces: From classical macroscopic descriptions to modern microscopic details ;Oxford University Press: Oxford, UK, 2004. [Google Scholar]
  11. MacFarlane D. R.; Forsyth M.; Izgorodina E. I.; Abbott A. P.; Annat G.; Fraser K. On the concept of ionicity in ionic liquids. Phys. Chem. Chem. Phys. 2009, 11, 4962–4967. 10.1039/b900201d. [DOI] [PubMed] [Google Scholar]
  12. Ueno K.; Tokuda H.; Watanabe M. Ionicity in ionic liquids: correlation with ionic structure and physicochemical properties. Phys. Chem. Chem. Phys. 2010, 12, 1649–1658. 10.1039/b921462n. [DOI] [PubMed] [Google Scholar]
  13. Stolwijk N. A.; Kösters J.; Wiencierz M.; Schönhoff M. On the extraction of ion association data and transference numbers from ionic diffusivity and conductivity data in polymer electrolytes. Electrochim. Acta 2013, 102, 451–458. 10.1016/j.electacta.2013.04.028. [DOI] [Google Scholar]
  14. Harris K. R. Relations between the fractional Stokes-Einstein and Nernst-Einstein Equations and velocity correlation coefficients in ionic liquids and molten salts. J. Phys. Chem. B 2010, 114, 9572–9577. 10.1021/jp102687r. [DOI] [PubMed] [Google Scholar]
  15. Kashyap H. K.; Annapureddy H. V. R.; Raineri F. O.; Margulis C. J. How is charge transport different in ionic liquids and electrolyte solutions?. J. Phys. Chem. B 2011, 115, 13212–13221. 10.1021/jp204182c. [DOI] [PubMed] [Google Scholar]
  16. Kirchner B.; Malberg F.; Firaha D. S.; Hollóczki O. Ion pairing in ionic liquids. J. Phys.: Condens. Matter 2015, 27, 463002. 10.1088/0953-8984/27/46/463002. [DOI] [PubMed] [Google Scholar]
  17. Shao Y.; Shigenobu K.; Watanabe M.; Zhang C. Role of viscosity in deviations from the Nernst–Einstein relation. J. Phys. Chem. B 2020, 124, 4774–4780. 10.1021/acs.jpcb.0c02544. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Hansen J.-P.; McDonald I. R.. Theory of simple liquids ;Academic Press: Amsterdam, Netherlands, 2006. [Google Scholar]
  19. Zhong E. C.; Friedman H. L. Self-diffusion and distinct diffusion of ions in solution. J. Phys. Chem. 1988, 92, 1685–1692. 10.1021/j100317a059. [DOI] [Google Scholar]
  20. Woolf L. A.; Harris K. R. Velocity correlation coefficients as an expression of particle-particle interactions in (electrolyte) solutions. J. Chem. Soc., Faraday Trans. 1 1978, 74, 933–947. 10.1039/f19787400933. [DOI] [Google Scholar]
  21. Vargas Barbosa N. M.; Roling B. Dynamic ion correlations in solid and liquid electrolytes: How do they affect charge and mass transport?. ChemElectroChem 2020, 7, 367–385. 10.1002/celc.201901627. [DOI] [Google Scholar]
  22. Fong K. D.; Self J.; McCloskey B. D.; Persson K. A. Ion correlations and their impact on transport in polymer-based electrolytes. Macromolecules 2021, 54, 2575–2591. 10.1021/acs.macromol.0c02545. [DOI] [Google Scholar]
  23. McDaniel J. G.; Son C. Y. Ion correlation and collective dynamics in BMIM/BF4-based organic electrolytes: from dilute solutions to the ionic liquid limit. J. Phys. Chem. B 2018, 122, 7154–7169. 10.1021/acs.jpcb.8b04886. [DOI] [PubMed] [Google Scholar]
  24. Li Z.; Bouchal R.; Mendez-Morales T.; Rollet A.-L.; Rizzi C.; Le Vot S.; Favier F.; Rotenberg B.; Borodin O.; Fontaine O.; et al. Transport properties of Li-TFSI water-in-salt Electrolytes. J. Phys. Chem. B 2019, 123, 10514–10521. 10.1021/acs.jpcb.9b08961. [DOI] [PubMed] [Google Scholar]
  25. Shao Y.; Hellström M.; Yllö A.; Mindemark J.; Hermansson K.; Behler J.; Zhang C. Temperature effects on the ionic conductivity in concentrated alkaline electrolyte solutions. Phys. Chem. Chem. Phys. 2020, 22, 10426–10430. 10.1039/C9CP06479F. [DOI] [PubMed] [Google Scholar]
  26. Fong K. D.; Self J.; McCloskey B. D.; Persson K. A. Onsager transport coefficients and transference numbers in polyelectrolyte solutions and polymerized ionic liquids. Macromolecules 2020, 53, 9503–9512. 10.1021/acs.macromol.0c02001. [DOI] [Google Scholar]
  27. Pfeifer S.; Ackermann F.; Sälzer F.; Schönhoff M.; Roling B. Quantification of cation-cation, anion-anion and cation-anion correlations in Li salt/glyme mixtures by combining very-low-frequency impedance spectroscopy with diffusion and electrophoretic NMR. Phys. Chem. Chem. Phys. 2021, 23, 628–640. 10.1039/D0CP06147F. [DOI] [PubMed] [Google Scholar]
  28. Park K.-H.; Choi S. R.; Choi J.-H.; Park S. a.; Cho M. Real-time probing of ion pairing dynamics with 2DIR spectroscopy. ChemPhysChem 2010, 11, 3632–3637. 10.1002/cphc.201000595. [DOI] [PubMed] [Google Scholar]
  29. Stange P.; Fumino K.; Ludwig R. Ion speciation of protic ionic liquids in water: Transition from contact to solvent-separated ion pairs. Angew. Chem., Int. Ed. 2013, 52, 2990–2994. 10.1002/anie.201209609. [DOI] [PubMed] [Google Scholar]
  30. Chaurasia S. K.; Singh R. K.; Chandra S. Ion-polymer complexation and ion-pair formation in a polymer electrolyte PEO:LiPF6 containing an ionic liquid having same anion: A Raman study. Vib. Spectrosc. 2013, 68, 190–195. 10.1016/j.vibspec.2013.08.001. [DOI] [Google Scholar]
  31. Mindemark J.; Lacey M. J.; Bowden T.; Brandell D. Beyond PEO–Alternative host materials for Li+ conducting solid polymer electrolytes. Prog. Polym. Sci. 2018, 81, 114–143. 10.1016/j.progpolymsci.2017.12.004. [DOI] [Google Scholar]
  32. Choo Y.; Halat D. M.; Villaluenga I.; Timachova K.; Balsara N. P. Diffusion and migration in polymer electrolytes. Prog. Polym. Sci. 2020, 103, 101220. 10.1016/j.progpolymsci.2020.101220. [DOI] [Google Scholar]
  33. Popovic J.; Brandell D.; Ohno S.; Hatzell K. B.; Zheng J.; Hu Y.-Y. Polymer-based hybrid battery electrolytes: theoretical insights, recent advances and challenges. J. Mater. Chem. A 2021, 9, 6050–6069. 10.1039/D0TA11679C. [DOI] [Google Scholar]
  34. Robinson R. A.; Stokes R. H.. Electrolyte solutions the measurement and interpretation of conductance, chemical potential, and diffusion in solutions of simple electrolytes ;Butterworths: London, UK, 1965. [Google Scholar]
  35. Luzar A. Resolving the hydrogen bond dynamics conundrum. J. Chem. Phys. 2000, 113, 10663–10675. 10.1063/1.1320826. [DOI] [Google Scholar]
  36. Rapaport D. Hydrogen bonds in water: Network organization and lifetimes. Mol. Phys. 1983, 50, 1151–1162. 10.1080/00268978300102931. [DOI] [Google Scholar]
  37. Luzar A.; Chandler D. Hydrogen-bond kinetics in liquid water. Nature 1996, 379, 55–57. 10.1038/379055a0. [DOI] [Google Scholar]
  38. Laage D.; Hynes J. T. On the residence time for water in a solute hydration shell: Application to aqueous halide solutions. J. Phys. Chem. B 2008, 112, 7697–7701. 10.1021/jp802033r. [DOI] [PubMed] [Google Scholar]
  39. Kirby B. J.; Jungwirth P. Charge scaling manifesto: A way of reconciling the inherently macroscopic and microscopic natures of molecular simulations. J. Phys. Chem. Lett. 2019, 10, 7531–7536. 10.1021/acs.jpclett.9b02652. [DOI] [PubMed] [Google Scholar]
  40. Wang J.; Wolf R. M.; Caldwell J. W.; Kollman P. A.; Case D. A. Development and testing of a general amber force field. J. Comput. Chem. 2004, 25, 1157–1174. 10.1002/jcc.20035. [DOI] [PubMed] [Google Scholar]
  41. Abraham M. J.; Murtola T.; Schulz R.; Páll S.; Smith J. C.; Hess B.; Lindahl E. GROMACS: High performance molecular simulations through multi-level parallelism from laptops to supercomputers. SoftwareX 2015, 1–2, 19–25. 10.1016/j.softx.2015.06.001. [DOI] [Google Scholar]
  42. Gudla H.; Zhang C.; Brandell D. Effects of solvent polarity on Li-ion diffusion in polymer electrolytes: An all-atom molecular dynamics study with charge scaling. J. Phys. Chem. B 2020, 124, 8124–8131. 10.1021/acs.jpcb.0c05108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Wheatle B. K.; Lynd N. A.; Ganesan V. Effect of polymer polarity on ion transport: A competition between ion aggregation and polymer segmental dynamics. ACS Macro Lett. 2018, 7, 1149–1154. 10.1021/acsmacrolett.8b00594. [DOI] [PubMed] [Google Scholar]
  44. Shen K.-H.; Hall L. M. Effects of ion size and dielectric constant on ion transport and transference number in polymer electrolytes. Macromolecules 2020, 53, 10086–10096. 10.1021/acs.macromol.0c02161. [DOI] [Google Scholar]
  45. Zhang Y.; Maginn E. J. Direct correlation between ionic liquid transport properties and ion pair lifetimes: A molecular dynamics study. J. Phys. Chem. Lett. 2015, 6, 700–705. 10.1021/acs.jpclett.5b00003. [DOI] [PubMed] [Google Scholar]
  46. Mogurampelly S.; Keith J. R.; Ganesan V. Mechanisms underlying ion transport in polymerized ionic liquids. J. Am. Chem. Soc. 2017, 139, 9511–9514. 10.1021/jacs.7b05579. [DOI] [PubMed] [Google Scholar]
  47. Shen K.-H.; Hall L. M. Ion conductivity and correlations in model salt-doped polymers: Effects of interaction strength and concentration. Macromolecules 2020, 53, 3655–3668. 10.1021/acs.macromol.0c00216. [DOI] [Google Scholar]
  48. Molinari N.; Mailoa J. P.; Kozinsky B. Effect of salt concentration on ion clustering and transport in polymer solid electrolytes: A molecular dynamics study of PEO-LiTFSI. Chem. Mater. 2018, 30, 6298–6306. 10.1021/acs.chemmater.8b01955. [DOI] [Google Scholar]
  49. Rey I.; Lassègues J.; Grondin J.; Servant L. Infrared and Raman study of the PEO-LiTFSI polymer electrolyte. Electrochim. Acta 1998, 43, 1505–1510. 10.1016/S0013-4686(97)10092-5. [DOI] [Google Scholar]
  50. Popovic J.; Pfaffenhuber C.; Melchior J. P.; Maier J. Determination of individual contributions to the ionic conduction in liquid electrolytes: Case study of LiTf/PEGDME-150. Electrochem. Commun. 2015, 60, 195–198. 10.1016/j.elecom.2015.09.009. [DOI] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

jz1c02474_si_001.pdf (574.9KB, pdf)

Articles from The Journal of Physical Chemistry Letters are provided here courtesy of American Chemical Society

RESOURCES