Oxygen Reduction Reaction in Nonaqueous Media

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Oxygen Reduction Reaction in Nonaqueous Media

EJ Calvo, University of Buenos Aires, Buenos Aires, Argentina


© 2018 Elsevier Inc. All rights reserved.

Introduction 831
Effect of Cations 832
Effects of Solvent 834
Stability of Solvent and Electrolyte 834
Influence of Water contamination 835
Conclusions 835
References 836
Further Reading 837

Introduction

The activation of molecular oxygen is important in diverse areas from respiration chain in biology to metal corrosion in materials
science and energy conversion in fuel cells and energy storage in metal–air batteries.1–4 Nonaqueous metal–air batteries using alkali
and alkaline earth metal anodes offer very high energy densities. The cathode active material, oxygen, need not be stored in the
battery and is supplied from the environment.5 Oxygen reduction in polar aprotic solvents is very different from aqueous solutions,
and strongly depends on the electrolyte solvent pair.6
The first electron-transfer step for the reduction of O2 in aprotic solvents containing hydrophobic cations such as alkyl ammo-
nium is a one-electron process to yield the radical anion, O2– with electron transfer to the p* oxygen orbital.7,8 Sawyer et al.9 re-
ported in 1966 that the electrochemical reduction of oxygen in dimethyl sulfoxide (DMSO) yields soluble superoxide:
O2 þ e/O
2 (1)

The standard reduction potential is Eo ¼  0.16 V in water but  0.60 V in DMF, which reflects very different solvation in both
solvents.
The name “superoxide” was coined for the potassium salt KO2 of the radical anion O2 $ in 1934 to differentiate to the combus-
tion product of alkali metals, oxides such as Na2O, K2O, and Li2O.10
The chemistry of superoxide has been extensively described by Sawyer.11–14 Ionic salts of superoxide generally formed from the
reaction of molecular oxygen with alkaline metals such as potassium, rubidium, or cesium are colored solids, yellow to orange.
These superoxides are paramagnetic with one unpaired electron per two oxygen atoms.
In 1969 two publications revealed the importance of superoxide in biology: the direct observation of superoxide by electron spin
resonance during an enzymatic reaction that involved dioxygen15 and the discovery of metalloproteins that catalyze the dispropor-
tionation of superoxide, that is, “superoxide dismutases” (SOD’S).16
Subsequent to electron transfer to O2, rapid protonation in the presence of water, methanol, etc. even in alkaline electrolyte the
lifetime of superoxide is very short due to the reaction:
HO$2 þ HO$2 /H2 O2 þ O2 (2)
1 1 
with a bimolecular rate constant k ¼ 8.6  10 M 5
s and a pKa(O2 ) ¼ 4.69. 17

Hþ þ O
2 #HO2 (3)

which implies superoxide is a weak acid.


However the reaction:
2O 
2 þ H2 O#O2 þ HO2 þ OH

(4)
has an equilibrium constant K ¼ 0.91  10 ; thus superoxide can promote proton transfer from substrates and solvents.
9

However in the presence of surfactants at pH 13 the lifetime of O2– could be as long as 1 min. In aprotic solvents, superoxide is
quite stable because disproportionation to give the peroxide dianion O22 is highly unfavorable. The effect of adding acidic
substrates to stable solutions of O2 in aprotic solvents has been determined by stopped-flow spectrophotometry and electro-
chemical measurements in dimethyl formamide (DMF) and in acetonitrile (MeCN).
A self-consistent mechanism for the proton-induced disproportion of O2 in aprotic media has been proposed for acidic
substrates:

HA þ O$2 #HO$2 þ A  (5)

831
832 Oxygen Reduction Reaction in Nonaqueous Media

following by Reaction (2).


The tendency of O2 to disproportionate via abstraction of protons from substrates and solvents is its most dominant reaction
characteristic. Superoxide radical anion is a powerful nucleophile in aprotic solvents; it does not exhibit such reactivity in water,
presumably because of its strong solvation in aqueous medium and its rapid hydrolysis and disproportionation.18

Effect of Cations

Afanasev and coworkers studied the absorption spectra of the superoxide formed by the electrochemical reduction of O2 in DMF
and acetonitrile and found that the superoxide ion exists in equilibrium with an ion pair (Bu4Nþ O2), in acetonitrile
(Keq ¼ 20 M 1), and as “free” solvated ion in DMF.19 However, the addition of the proton donors HX (water or ethanol) to O2–
solutions in DMF and acetonitrile caused the formation of new ion pairs ðBu4 Nþ O2 $ Þ2HX.
Unlike the reversible behavior of O2/O2– in electrolytes of aprotic solvents with large cations such as Bu4Nþ, in alkali metal hexa-
fluorophosphate or perchlorate an irreversible electrochemical wave has been reported.20,21 Large cations like TBAþ interact weakly
with O2, resulting in soluble ion pair but small cations are strong Lewis acids that interact tightly with O2– and result in unstable
LiO2.
A detailed study of the oxygen reduction reaction (ORR) in nonaqueous electrolytes in relation to the lithium-air battery has
been reported by the group of Abraham.20–24 In lithium containing electrolyte the cyclic voltammetry exhibits an irreversible reduc-
tion peak with formation of a one-electron reduction product, LiO2, at 2.5 V versus Li/Liþ in DMSO with passivation of the elec-
trode surface. Lithium superoxide can subsequently undergo disproportionation into insoluble lithium peroxide, Li2O2 and O2, or
a second electron transfer to form lithium peroxide according to the following reactions.
O2 þ Liþ þ e/LiO2 E0 ¼ 3:0 V (6)

2LiO2 /Li2 O2 þ O2 (7)

LiO2 þ Liþ þ e/Li2 O2 E0 ¼ 3:1 V (8)


Further electron transfer at higher overpotential may lead to formation of lithium oxide:
Li2 O2 þ 2Liþ þ 2e/2Li2 O (9)
In the absence of lithium ions in different solvents or at low lithium ion concentration in solvents which strongly solvate lithium
cations the back sweep shows oxidation of the stable soluble superoxide ion which can also be detected with a rotating ring disc
electrode (RRDE) at a Au ring under convective-diffusion conditions.25–27
The formation of the O2 reduction products in Eqs. (7)–(9) has been demonstrated spectroscopically by surface enhanced
Raman spectroscopy (SERS) for AueO adsorbed O2– at 491 cm 1 and OeO stretch of adsorbed O2– at 1109 cm 1 and OeO stretch
at 808 cm 1 for Li2O2.28–30
In lithium ion concentrations above the solubility of oxygen, the back scan in cyclic voltammetry shows only oxidation current
due to oxygen evolution reaction at high overpotentials:
Li2 O2 /O2 þ 2Liþ þ 2e (10)
Fig. 1 depicts the reversible reduction of O2 in 0.1 M TBAPF6 in DMSO (upper panel) and in 2 mM LiPF6 þ 98 mM TBAPF6
(middle panel), and in 5 mM LiPF6 þ 95 mM TBAPF6 (lower panel) on Au at 20 mV s 1 and simultaneous electrochemical quartz
crystal microbalance (EQCM) The solubility of O2 in DMSO containing lithium is 2 mM; for Liþ concentration above this value
(5 mM) a new cathodic peak at 2.0–2.5 V is observed in addition to the reversible O2/O2– which is attributed to the formation
of Li2O2 on the surface, as reported for increasing concentrations of lithium ions.28,29 We also observe mass increase with the
EQCM, passivation of the surface, and disappearance of the superoxide oxidation peak in the back scan for Liþ concentrations larger
than 5 mM. A new peak at 3 V corresponds to the oxidation of surface Li2O2 as has been demonstrated by differential electrochem-
ical mass spectrometry31 and Raman spectroscopy.29
A systematic study of small alkali metal cations, Liþ, Naþ, and Kþ, together with large cations Bu4Nþ and cations of ionic liquids
(IL) by the group of Abraham21 has shed light into the cation size on the Lewis acidity of the cations and the stability of O2 reduc-
tion products such as O2. The relative Lewis acidity of the different cations was measured by the 13C NMR chemical shift and spin
relaxation time (T1) of the C]O functional group in propylene carbonate due to its interaction with each of the cations by forma-
tion of acid–base complexes. The Liþ (ionic radius 60 pm) with relative high positive charge density or strong Lewis acidity formed
the strongest complex with C]O with a large NMR chemical shift while very large TBAþ(494 pm) and IL cations (239–330 pm)
with low charge densities exhibited the lowest Lewis acidity and thus the least chemical shift. The large cations are soft acids while
Kþ, Naþ, and LIþ are increasingly harder acids with Liþ being the strongest Lewis acid. The organic solvents of the nonaqueous elec-
trolytes are Lewis bases characterized by their Gutmann donor number (DN).32 Therefore, by replacing large TBAþ with Liþ, Naþ, or
Kþ the reversible nature of the ORR in aprotic solvents is severely suppressed. Clearly, O2– is stabilized as TBA–O2 in tertaalkyl-
ammonium salt solutions.
Oxygen Reduction Reaction in Nonaqueous Media 833

Fig. 1 Cyclic voltammetry of O2 in 0.1 M TBAPF6 in DMSO (upper panel), in 2 mM LiPF6 þ 98 mM TBAPF6 (middle panel), and in 5 mM
LiPF6 þ 95 mM TBAPF6 (lower panel) on Au at 20 mV s 1 and simultaneous EQCM gravimetry. Taken from del Pozo M.; Torres, W.R.; Herrera, S.E.;
Calvo, E.J. New Evidence of LiO2 Dismutation in Lithium–Air Battery Cathodes. ChemElectroChem 2016, 3(10): 1537–1540.

Increasing the cation size from Liþ to Kþ alters the cyclic voltammetry. Potassium with a large ionic radius (220 pm) at 0.1 M
KPF6 has a similar cyclic voltammetry as lithium and sodium; however at 1 M KPF6 it is similar to that of TBA.20
In pure IL the electroreduction of molecular oxygen also results in a reversible wave as shown in Fig. 2 for 1-ethyl-3-methyl-imi-
dazolium bis(trifluoromethylsulfonyl) imide (EMITFSI).33 However, addition of small alkali metal cations affects the reversibility
of the ORR which is intermediate between those in Li and TBA salt solutions.
Fig. 2 shows a good correlation between the relative size of the cation (charge density) and the ORR mechanism as well as the
stability of the ORR intermediates and the final products formed. The effect of the cation Lewis acidity and the donor properties of
the solvent on the relative stability of superoxide have been explained by Abraham and coworkers by the Hard Soft Acid Base
(HSAB) theory of ORR in nonaqueous electrolytes.
Pearson’s HSAB theory34 states that hard acids prefer to be associated with hard bases and soft bases prefer soft acids.

Fig. 2 Cyclic Voltammogram of neat EMITFSI along with various salts at 25 mM concentration on a GC electrode at 0.1 V s 1. Taken from Allen,
C.J., et al. Oxygen Reduction Reactions in Ionic Liquids and the Formulation of a General ORR Mechanism for Li–Air Batteries. J. Phys. Chem. C
2012, 116(39), 20755–20764.
834 Oxygen Reduction Reaction in Nonaqueous Media

Effects of Solvent

Abraham and coworkers have suggested that the stability of Liþ–O2– ion pair increases with solvent DN, which is a measure of the
solvation enthalpy of the Lewis acid SbCl5 taken as reference.32 The DN for several solvents follows the order, acetonitrile (14.1) <
tetraethylene glycol-dimethyl ether (TEGDME) (16.6) < (DME) (20.0) < (DMSO) (29.8). The work of Bruce et al.28 has shown that
increasing the solvent DN leads to increase Liþ–O2– solubility and capacity of LieO2 cathodes upon discharge.
Therefore, the determining factor for the overall ORR mechanism and products formed is the net Lewis acidity of the solvated
Liþ(solvent)n, with n typically 4.35 In low DN acetonitrile no soluble O2– can be detected by RRDE, while in high DN DMSO soluble
superoxide is stable. Due to preferential solvation of Liþ by DMSO even at 0.1 M DMSO in acetonitrile superoxide could be detected
by RRDE.36
This leads to two different mechanisms for the ORR in Liþ-containing nonaqueous solvents as has recently been proposed,
depending on the growth of Li2O2 at the cathode surface: (a) surface reaction or (b) solution-phase reaction, as compared in
Scheme 1.37–39 The ORR surface mechanism is operative in acetonitrile while a solution-phase mechanism takes place in DMSO.
A recent calculation by the group of Shao-Horn has shown that the greater Liþ solvation lowers the Liþ/Li potential while greater
O2– solvation increases the of O2/O2– potential.40 Thus, the O2/TBAþ–O2– redox potential referred to the Liþ/Li scale increases with
combined solvation of superoxide and lithium ion. This can be seen in Fig. 3 for O2 saturated PYR14TFSIþ at different LiTFSI
concentrations, 0–50 mM. The formation of Liþ–O2– ion pair takes place at more positive potentials than that of the PYR14þ–O2–
detected at the RRDE.41
In contrast to LieO2 batteries, nonaqueous NaeO2 produces sodium superoxide, NaO2 in “dry” diglyme-based electrolyte and
has attracted attention since it exhibits more reversible electrochemistry than their lithium counterparts and sodium is very abun-
dant in nature.42
Naþ þ O2 þ e/NaO2 E0 ¼ 2:27 V (11)

NaO2 þ O2 þ e/Li2 O2 E0 ¼ 2:33 V (12)

Stability of Solvent and Electrolyte

In spite of the very high theoretical energy densities of 3500 W h kg 1, nonaqueous lithium–air batteries (LieO2 cells) two prob-
lems still remain for their practical application, namely the battery capacity fading and high recharge overpotentials. These are due
to the chemical instability of solvent, electrolyte, and carbon electrodes in the presence of kinetically highly reactive ORR interme-
diates and products, LiO2 and Li2O2. Superoxide radical anion is very reactive as nucleophile and strong base; it attracts solvents and

Scheme 1 Scheme of reactions for surface and solution-phase mechanisms of the ORR in lithium-containing electrolyte. Taken from del Pozo M.;
Torres, W.R.; Herrera, S.E.; Calvo, E.J. New Evidence of LiO2 Dismutation in Lithium–Air Battery Cathodes. ChemElectroChem 2016, 3(10): 1537–
1540.
Oxygen Reduction Reaction in Nonaqueous Media 835

Fig. 3 Cyclic voltammetry experiment of PYR14TFSIþ O2 saturated at different LiTFSI concentrations 0–50 mM at 50 mV s 1. Taken from Mohzhu-
khina, N.; Tesio, A.Y.; del Pozo, M.; Calvo, E.J. Lithium Ion Concentration Effect in PYR14TFSI Ionic Liquid for Li–O2 Battery Cathodes. J. Electrochem.
Soc. 2017, 164(8), H5277–H5279.

electrolytes by hydrogen abstraction. Li2O2 is also a strong oxidant in contact with the electrolyte and carbon electrodes. These side
reactions were neglected in the early stages of LieO2 batteries studies.43
A number of solvents have been considered, including alkyl carbonates, dimethoxyethane (DME), glymes, DMSO DMF, IL, etc.
Unfortunately, all of them are degraded by the ORR reactive products.
A very large effort has been reported in recent years using a large variety of in situ and ex situ analytical techniques to study the
spurious reactions leading to capacity fading a high recharge overpotential. These have included XRD, XPS, AFM, Raman, FTIR, TEM,
NMR, and DEMS. Formation of lithium carbonate, LiF, dimethyl sulfone, and oxidation of IL cations are some of the unwanted side
reaction products detected.41,44–46
A recent report has suggested singlet oxygen 1O2 generation as a major cause for parasitic reactions during cycling of aprotic
lithium–oxygen batteries.47

Influence of Water contamination

Traces of water are unavoidable in metal air batteries “breathing air” from the atmosphere. Therefore the effect of traces of water on
the chemistry of superoxide has shown that increasing the amount of water from 10 mM to 3 M a decrease in the O2– oxidation peak
is observed due to the superoxide protonation reaction followed by disproportionation forming hydrogen peroxide. A possible
mechanism is21:

TBA þ O þ þ $
2 þ H3 O /TBA þ HO2 þ H2 O (13)

followed by Reaction (2).


A hydrogen bonded complex with its positive dipole oriented toward the electrode can facilitate the electron transfer acting as
a redox mediator and decreasing the redox potential of the O2/O2– couple in the presence of water.

Conclusions

The first electron transfer to oxygen in nonaqueous media with large cations is the formation of the radical anion superoxide by an
electrochemical one-electron reaction. In the presence of small alkali cations (Liþ, Naþ, Kþ) the ORR is irreversible and ion–ion and
ion–solvent interactions determine the reaction mechanism and products. Highly reactive lithium and sodium superoxide and
peroxides can degrade solvent, electrolyte, and carbon electrodes with metal ion battery capacity fading and high recharge overpo-
tential. Traces of water may produce hydrogen peroxide and affect the mechanism.

See also: Bioinspired Electrocatalysis for the Oxygen Reduction Reaction; Graphene-Based Nanostructures in Electrocatalytic Oxygen Reduction;
Oxygen Reduction on Platinum Single Crystal Electrodes.
836 Oxygen Reduction Reaction in Nonaqueous Media

References

1. Calvo, E. J. The Kinetics of Oxygen Electroreduction: A Long Way From Iron Rust to Lithium–Air Batteries. Mater. Corros. 2014, 65 (4), 345–350.
2. Gewirth, A. A.; Thorum, M. S. Electroreduction of Dioxygen for Fuel-Cell Applications: Materials and Challenges. Inorg. Chem. 2010, 49 (8), 3557–3566.
3. Kjaergaard, C. H.; Rossmeisl, J.; Nørskov, J. K. Enzymatic Versus Inorganic Oxygen Reduction Catalysts: Comparison of the Energy Levels in a Free-Energy Scheme. Inorg.
Chem. 2010, 49 (8), 3567–3572.
4. Markovic, N. M.; Schmidt, T. J.; Stamenkovic, V.; Ross, P. N. Oxygen Reduction Reaction on Pt and Pt Bimetallic Surfaces: A Selective Review. Fuel Cells 2001, 1 (2),
105–116.
5. Abraham, K. M. A Brief History of Non-Aqueous Metal–Air Batteries. ECS Trans. 2008, 3, 67–71.
6. Sharon, D.; et al. Aprotic Metal-Oxygen Batteries: Recent Findings and Insights. J. Solid State Electrochem. 2017, 21 (7), 1861–1878.
7. Maricle, D. L.; Hodgson, W. G. Reduction of Oxygen to Superoxide Anion in Aprotic Solvents. Anal. Chem. 1965, 37 (12), 1562–1565.
8. Peover, M. E.; White, B. S. Electrolytic Reduction of Oxygen in Aprotic Solvents: The Superoxide Ion. Electrochim. Acta 1966, 11 (8), 1061–1067.
9. Sawyer, D. T.; Roberts, J. L., Jr. Electrochemistry of Oxygen and Superoxide Ion in Dimethylsulfoxide at Platinum, Gold and Mercury Electrodes. J. Electroanal. Chem. 1966,
12 (2), 90–101.
10. Neuman, E. W. Potassium Superoxide and the Three-Electron Bond. J. Chem. Phys. 1934, 2, 31.
11. Sawyer, D. T.; Calderwood, T. S.; Yamaguchi, K.; Angelis, C. T. Synthesis and Characterization of Tetramethylammonium Superoxide. Inorg. Chem. 1983, 22 (18),
2577–2583.
12. Sawyer, D. T.; Chiericato, G.; Angelis, C. T.; Nanni, E. J.; Tsuchiya, T. Effects of Media and Electrode Materials on the Electrochemical Reduction of Dioxygen. Anal. Chem.
1982, 54 (11), 1720–1724.
13. Sawyer, D. T.; Gibian, M. J. The Chemistry of Superoxide Ion. Tetrahedron 1979, 35 (12), 1471–1481.
14. Sawyer, D. T.; Gibian, M. J.; Morrison, M. M.; Seo, E. T. On the Chemical Reactivity of Superoxide Ion [6]. J. Am. Chem. Soc. 1978, 100 (2), 627–628.
15. Knowles, P. F.; Gibson, J. F.; Pick, F. M.; Bray, R. C. Electron-Spin-Resonance Evidence for Enzymic Reduction of Oxygen to a Free Radical, the Superoxide Ion. Biochem. J.
1969, 111 (1), 53–58.
16. McCord, J. M.; Fridovich, I. Superoxide Dismutase. An Enzymic Function for Erythrocuprein (Hemocuprein). J. Biol. Chem. 1969, 244 (22), 6049–6055.
17. Sawyer, D. T.; Valentine, J. S. How Super Is Superoxide? Acc. Chem. Res. 1981, 14 (12), 393–400.
18. Wilshire, J.; Sawyer, D. T. Redox Chemistry of Dioxygen Species. Acc. Chem. Res. 1979, 2 (3), 105–110.
19. Afanas’ev, I. B.; Kuprianova, N. S.; Polozova, N. I. Kinetics and Mechanism of the Reactions of the Superoxide Ion in Solutions. I. Absorption Spectra and Interaction With
Solvents and Cations. In. J. Chem. Kinet. 1983, 15, 1045–1056.
20. Laoire, C. O.; Mukerjee, S.; Abraham, K. M.; Plichta, E. J.; Hendrickson, M. A. Elucidating the Mechanism of Oxygen Reduction for Lithium–Air Battery Applications. J. Phys.
Chem. C 2009, 113 (46), 20127–20134.
21. Abraham, K. M. Electrolyte-Directed Reactions of the Oxygen Electrode in Lithium–Air Batteries. J. Electrochem. Soc. 2015, 162 (2), A3021–A3031.
22. Abraham, K. M. Lithium–Air and Other Batteries beyond Lithium–Ion Batteries. In Lithium Batteries, John Wiley & Sons, Inc.: Wiley Electrochemical Society, Pennington, NJ,
USA, 2013; pp 161–190. ISBN: 978-1-118-18365-6.
23. Laoire, C. O.; Mukerjee, S.; Abraham, K. M.; Plichta, E. J.; Hendrickson, M. A. Influence of Nonaqueous Solvents on the Electrochemistry of Oxygen in the Rechargeable
Lithium  Air Battery. J. Phys. Chem. C 2010, 114 (19), 9178–9186.
24. Laoire, C. O.; Mukerjee, S.; Plichta, E. J.; Hendrickson, M. A.; Abraham, K. M. Rechargeable Lithium/TEGDME-LiPF[Sub 6]/O[Sub 2] Battery. J. Electrochem. Soc. 2011,
158 (3), A302.
25. Herranz, J.; Garsuch, A.; Gasteiger, H. A. Using Rotating Ring Disc Electrode Voltammetry to Quantify the Superoxide Radical Stability of Aprotic Li–Air Battery Electrolytes.
J. Phys. Chem. C 2012, 116 (36), 19084–19094.
26. Torres, W.; Mozhzhukhina, N.; Tesio, A. Y.; Calvo, E. J. A Rotating Ring Disk Electrode Study of the Oxygen Reduction Reaction in Lithium Containing Dimethyl Sulfoxide
Electrolyte: Role of Superoxide. J. Electrochem. Soc. 2014, 161 (14), A2204–A2209.
27. Trahan, M. J.; Mukerjee, S.; Plichta, E. J.; Hendrickson, M. A.; Abraham, K. M. Studies of Li–Air Cells Utilizing Dimethyl Sulfoxide-Based Electrolyte. J. Electrochem. Soc.
2012, 160 (2), A259–A267.
28. Johnson, L.; et al. The Role of LiO2 Solubility in O-2 Reduction in Aprotic Solvents and its Consequences for Li-O-2 Batteries. Nat. Chem. 2014, 6 (12), 1091–1099.
29. Yu, Q.; Ye, S. In Situ Study of Oxygen Reduction in Dimethyl Sulfoxide (DMSO) Solution: A Fundamental Study for Development of the Lithium–Oxygen Battery. J. Phys. Chem.
C 2015, 119 (22), 12236–12250.
30. Gunasekara, I.; Mukerjee, S.; Plichta, E. J.; Hendrickson, M. A.; Abraham, K. M. Microelectrode Diagnostics of Lithium–Air Batteries. J. Electrochem. Soc. 2014, 161 (3),
A381–A392.
31. Luntz, A. C.; McCloskey, B. D. Nonaqueous Li–Air Batteries: A Status Report. Chem. Rev. 2014, 114 (23), 11721–11750.
32. Gutmann, V. Empirical Parameters for Donor and Acceptor Properties of Solvents. Electrochim. Acta 1976, 21 (9), 661–670.
33. Allen, C. J.; et al. Oxygen Reduction Reactions in Ionic Liquids and the Formulation of a General ORR Mechanism for Li–Air Batteries. J. Phys. Chem. C 2012, 116 (39),
20755–20764.
34. Pearson, R. G. Hard and Soft Acids and Bases. J. Am. Chem. Soc. 1963, 85 (22), 3533–3539.
35. Semino, R.; Zaldívar, G.; Calvo, E. J.; Laria, D. Lithium Solvation in Dimethyl Sulfoxide–Acetonitrile Mixtures. J. Chem. Phys. 2014, 141 (21), 214509.
36. Calvo, E. J.; Mozhzhukhina, N. A Rotating Ring Disk Electrode Study of the Oxygen Reduction Reaction in Lithium Containing Non Aqueous Electrolyte. Electrochem. Commun.
2013, 31, 56–58.
37. Aetukuri, N. B.; et al. Solvating Additives Drive Solution-Mediated Electrochemistry and Enhance Toroid Growth in Non-Aqueous Li–O(2) Batteries. Nat. Chem. 2015, 7 (1),
50–56.
38. McCloskey, B. D.; Bethune, D. S.; Shelby, R. M.; Girishkumar, G.; Luntz, A. C. Solvents’ Critical Role in Nonaqueous Lithium–Oxygen Battery Electrochemistry. J. Phys. Chem.
Lett. 2011, 2 (10), 1161–1166.
39. Adams, B. D.; et al. Current Density Dependence of Peroxide Formation in the Li–O2 Battery and Its Effect on Charge. Energy Environ. Sci. 2013, 6 (6), 1772.
40. Kwabi, D. G.; et al. Experimental and Computational Analysis of the Solvent-Dependent O2/Li(þ)O2() Redox Couple: Standard Potentials, Coupling Strength, and
Implications for Lithium–Oxygen Batteries. Angewandte Chemie - International Edition 2016, 55 (9), 3129–3134.
41. Mohzhukhina, N.; Tesio, A. Y.; del Pozo, M.; Calvo, E. J. Lithium Ion Concentration Effect in PYR14TFSI Ionic Liquid for Li–O2 Battery Cathodes. J. Electrochem. Soc. 2017,
164 (8), H5277–H5279.
42. Hartmann, P.; et al. A Comprehensive Study on the Cell Chemistry of the Sodium Superoxide (NaO2) Battery. Phys. Chem. Chem. Phys. 2013, 15 (28), 11661–11672.
43. Sharon, D.; et al. Lithium–Oxygen Electrochemistry in Non-Aqueous Solutions. Isr. J. Chem. 2015, 55 (5), 508–520.
44. Marchini, F.; et al. Surface Study of Lithium–Air Battery Oxygen Cathodes in Different Solvent-Electrolyte Pairs. Langmuir 2015, 31 (33), 9236–9245.
45. Mozhzhukhina, N.; Méndez De Leo, L. P.; Calvo, E. J. Infrared Spectroscopy Studies on Stability of Dimethyl Sulfoxide for Application in a Li–Air Battery. J. Phys. Chem. C
2013, 117 (36), 18375–18380.
46. Amin, H. M. A.; Molls, C.; Bawol, P. P.; Baltruschat, H. The Impact of Solvent Properties on the Performance of Oxygen Reduction and Evolution in Mixed Tetraglyme-Dimethyl
Sulfoxide Electrolytes for Li–O2 Batteries: Mechanism and Stability. Electrochim. Acta 2017, 245, 967–980.
Oxygen Reduction Reaction in Nonaqueous Media 837

47. Mahne, N.; et al. Singlet Oxygen Generation as a Major Cause for Parasitic Reactions during Cycling of Aprotic Lithium–Oxygen Batteries. Nat. Energy 2017, 2,
17036.

Further Reading

del Pozo, M.; Torres, W. R.; Herrera, S. E.; Calvo, E. J. New Evidence of LiO2 Dismutation in Lithium–Air Battery Cathodes. ChemElectroChem 2016, 3 (10), 1537–1540.

You might also like