Monitoring the Electrochemical Processes in the Lithium–Air Battery by Solid State NMR Spectroscopy (2024)

As a library, NLM provides access to scientific literature. Inclusion in an NLM database does not imply endorsem*nt of, or agreement with, the contents by NLM or the National Institutes of Health.
Learn more: PMC Disclaimer | PMC Copyright Notice

Monitoringthe Electrochemical Processes in the Lithium–AirBattery by Solid State NMR Spectroscopy (1)

J Phys Chem C Nanomater Interfaces. 2013 Dec 27; 117(51): 26929–26939.

Published online 2013 Nov 27. doi:10.1021/jp410429k

PMCID: PMC3905693

PMID: 24489976

Author information Article notes Copyright and License information PMC Disclaimer

Associated Data

Supplementary Materials

Abstract

Monitoringthe Electrochemical Processes in the Lithium–AirBattery by Solid State NMR Spectroscopy (2)

A multi-nuclearsolid-state NMR approach is employed to investigatethe lithium–air battery, to monitor the evolution of the electrochemicalproducts formed during cycling, and to gain insight into processesaffecting capacity fading. While lithium peroxide is identified by 17O solid state NMR (ssNMR) as the predominant product in thefirst discharge in 1,2-dimethoxyethane (DME) based electrolytes, itreacts with the carbon cathode surface to form carbonate during thecharging process. 13C ssNMR provides evidence for carbonateformation on the surface of the carbon cathode, the carbonate beingremoved at high charging voltages in the first cycle, but accumulatingin later cycles. Small amounts of lithium hydroxide and formate arealso detected in discharged cathodes and while the hydroxide formationis reversible, the formate persists and accumulates in the cathodeupon further cycling. The results indicate that the rechargeabilityof the battery is limited by both the electrolyte and the carbon cathodestability. The utility of ssNMR spectroscopy in directly detectingproduct formation and decomposition within the battery is demonstrated,a necessary step in the assessment of new electrolytes, catalysts,and cathode materials for the development of a viable lithium–oxygenbattery.

1. Introduction

Thecontinued increase in global energy consumption and the shifttoward electrification of transportation call for significant improvementsin current lithium ion battery technology. Such improvements requirethe design of new materials and new chemistries to enable the developmentof energy storage devices offering higher energy densities. The lithiumair battery is considered a promising candidate for such applications,as it can potentially deliver an order of magnitude higher gravimetricenergy density than conventional lithium ion batteries. This highenergy is based on the reversible reaction between lithium and oxygen,oxygen being provided from the atmosphere, forming lithium peroxide.1

Despite initial results demonstrating reversiblecycling of thelithium–oxygen cell using an aprotic electrolyte with capacitiesof more than 1000 mAh/g,2 there are severalchallenges facing the successful development of this battery. Amongthese are the identification of stable electrolyte salts and solvents,development of inert, porous, and conductive cathode materials, aswell as design of catalytic species for reducing the overpotentialsof both the discharge and charge processes.35 Several studieshave demonstrated and discussed the issue of electrolyte stabilityin the presence of the highly reactive superoxide species formed duringthe discharge process. The use of common carbonate solvents was shownto lead to the formation of various electrolyte decomposition products,mostly lithium carbonate.69 Ether solvents, although considered relatively stablein the initial cycles, were shown to decompose upon extended cycling.10 However, as their use enables the formationof significant amounts of peroxide, tetraethyleneglycol (TEGDME) and1,2-dimethoxyethane (DME) are used in many studies and are often consideredto be relatively inert.1113 Other aprotic solvents examinedinclude dimethyl sulfoxide (DMSO) and dimethyl formamide (DMF), whichwere also shown to decompose to a certain extent.1417 Similarly, the stability of theelectrolyte salts have been investigated. Studies employing solutionnuclear magnetic resonance (NMR) spectroscopy, X-ray diffraction (XRD)and X-ray photoelectron spectroscopy (XPS) have identified that theinstability of the lithium salt can reduce the cycle life of the celldue to side reactions that depend strongly on the combination of saltand solvent.1821 Much attention has also been given to the porous carbon cathodein the cell. While initially it was thought that the role of the cathodeis to mediate the reaction between oxygen and lithium by allowingelectron conduction to the site of reaction and housing the reactionproducts within its pores, recent studies have shown that the carboncathode itself can affect the morphology and mechanism of the desiredreaction product, lithium peroxide.22 Furthermore,it was suggested that carbonaceous electrodes may not be stable inall voltage windows and may be prone to decomposition.16,2325

The carbon stability issues are exacerbatedby the large over potentials,and thus higher potentials, associated with the charging process.The charging voltage can, in principle, be lowered by the additionof a catalyst to the electrode formulation and numerous precious metaland metal oxides particles have been proposed as potential catalysts.2629 Despite the apparent decrease in charging potential obtained withsome of these species, it is becoming increasingly apparent that manyof them lead to additional electrolyte decomposition. Furthermore,it remains unclear whether catalysis can take place at all in a systemwith limited or no solubility of the reaction products.30

Clearly the development of a viable lithium-airbattery is stillat the stage where insight into the electrochemical reactions takingplace at the electrode is a necessity. The interplay between the electrodematerial, the reactants lithium and oxygen, and their products aswell as the electrolyte solution must be fully understood so thatrobust materials can be designed. While new carbon structures, compositeelectrodes, catalysts, and electrolyte candidates are constantly introduced,their viability and role must be assessed by a careful characterizationof the electrochemical processes that they give rise to.

NMRspectroscopy can be particularly enlightening as it allowsthe detection of the entire bulk of the cathode, detecting both amorphousand ordered phases and provides a thorough description of the electrochemicalprocesses taking place. This contrasts with other analytical toolssuch as X-ray diffraction, differential mass spectroscopy (DEMS) andsurface techniques such as XPS, which focus on smaller fractions ofthe electrochemical processes: crystalline, gaseous, or surface componentsrespectively.

Although lithium peroxide was shown to form byvarious techniquessuch as XRD, secondary emission spectroscopy (SEM), and infrared (IR),8,25,31,32 its quantification via these methods is not trivial. We have recentlyshown the strength of 17O ssNMR in investigating this systemin part due to the unique NMR parameters of lithium peroxide,33 demonstrating that the main discharge productin DME is indeed lithium peroxide. However, several other resonanceswere observed in the 17O spectra of discharged electrodesthat were assigned to a series of decomposition products. Here weextend the NMR approach further: by assembling a library of possiblereaction products and their NMR signatures we can clearly detect andquantify the amount of lithium peroxide and electrolyte decompositionproducts formed in working cells.

Since most of the reactionproducts formed electrochemically containO and Li, it should be possible to identify them via the acquisitionof their 6,7Li and 17O NMR spectra. Althoughlithium NMR spectra are straightforward to acquire, due to the ionicnature of Li+ in the (diamagnetic) compounds studied here,the variations between electronic environments in the various compoundsare not sufficiently large to cause pronounced differences in chemicalshifts.34 In particular, the lithium peroxideand carbonate chemical shifts differ by only 0.18 ppm35,36 and hence they are difficult to resolve even when using the lowerabundant 6Li nucleus (which generally gives rise to sharpersignals due to the weaker hom*onuclear couplings). 17O,which has a spin 5/2, is a much more sensitive probe to its chemicalenvironment and is therefore the main probe nucleus in this study.The quadrupole coupling constants, Cq,are often of the order of few megahertz, resulting in distinct secondorder quadrupole line shapes, which, in combination with the chemicalshift, allows the identification of most of the electrochemical productsformed in the cell.

By performing a multinuclear study of 17O, 6,7Li, 1H, and 13C ssNMR,we identify and quantifythe different nonsoluble products formed in the pores and on the surfaceof the carbon electrode. Furthermore, we monitor the evolution ofthese products as a function of state of charge and discuss the resultsin the context of prior mechanistic studies of electrolyte decomposition.Specific carbon-13 labeling of the electrode material and of the electrolyteis used to identify decomposition pathways, and correlation spectroscopyis used to spatially locate the formed species. We expect this approachto be beneficial in evaluating new electrode and electrolytes forlithium–air batteries.

2. Experimental Section

a. Electrode Fabrication

Oxygen electrodeswere prepared from a mixture of 24 wt % superP Li carbon (Timcal),38 wt % polyvinylidene fluoride (PVDF) binder, and 38 wt % dibutylphthalate(DBP, Sigma-Aldrich) in acetone. The slurry was then spread into aself-supporting film of thickness 150 μm, cut into disc shapewith 1/2″ diameter, and washed with diethyl ether to removethe DBP. The final film contained 40 wt % carbon. The electrodes werethen vacuum-dried at 115 °C and taken into an argon filled gloveboxwithout exposure to ambient atmosphere.

13C-enrichedelectrodes were made following the same procedure using 13C enriched (99%) amorphous carbon powder from Cambridge Isotope Laboratories.

b. Cell Assembly and Electrochemistry

Lithium–oxygencells were assembled in an argon-filled gloveboxusing standard 3205 coin cells (Hohsen Corp.) that were modified with1 mm diameter holes punched to their top case. Cells were assembledby stacking a disc of lithium foil (0.38 mm thickness, Sigma-Aldrich)on top of a stainless steel current collector, followed by a borosilicateglass fiber separator (Whatman). Electrolyte made of 1 M lithium bis(trifluoromethanesulfonyl)imide(LiTFSI, Sigma Aldrich, vacuum-dried at 250 °C prior to use for12 h) in 1,2-dimethoxyethane (DME, Sigma Aldrich, vacuum distilledand stored over molecular sieves in an argon glovebox) was added tothe separator. The self-supporting carbon electrode was then placedon top, followed by a stainless steel mesh (Advert Materials) andthe punched coin cell top. The cell was pressed closed, excess electrolyteremoved, and sealed in a glass chamber with two Young valves. Followingassembly in the glovebox, cells were flushed with oxygen gas for 30min through the valves. After adding the oxygen, the cells were restedfor 10–20 h and then cycled on an Arbin battery cycler.

Following cycling, cells were disassembled in the glovebox and thecathodes were extracted, washed with anhydrous acetonitrile, and vacuum-driedovernight in the glovebox’s prechamber and packed into theNMR rotors without exposure to ambient atmosphere.

For 6Li measurements, 6Li-enriched lithiumfoil (Cambridge Isotope Laboratories) was used as the anode.

c. 17O Isotope Enrichment

For 17O NMR measurements the cell assembly was performed as describedabove. Following the addition of natural abundance oxygen, the cellwas connected to a vacuum line. 17O enriched oxygen gas(60–70% enrichment, Isotec and Cambridge Isotope Laboratories)was connected to the line. The pressure in the cell was reduced toabout 0.8 atm and refilled back to 1 atm with the 17O enrichedoxygen gas resulting in about 20–25% 17O enrichedoxygen gas. The cell was then rested for 10–20 h and cycledas above.

d. Synthesis of Model Compounds

17O-enriched LiOH was synthesized following the procedure of Abys etal.37 from the reaction between n-butyllithium in hexane (52 mL) and 17O enriched (10%) water(1.5 mL) in 100 mL dry tetrahydrofuran, resulting in 1.5 g of LiOH. 17O-enriched Li2O was obtained by heating the aboveLiOH to 700 °C under vacuum, raising the temperature in severalsteps, again following the procedure by Abys et al.3717O enriched Li2CO3 wasmade by placing the above LiOH in a tube furnace heated to 100 °Cand flowing CO2 gas at a rate of 15 cm3/minfor 12 h. The formation of the phases and their purity was confirmedby X-ray powder diffraction collected on a Panalytical X’PertPro diffractometer.

e. Solid State NMR Spectroscopy

1H ssNMR measurements were performed on a Bruker 700MHz Avance IIIspectrometer using a 1.3 mm double resonance probe. A rotor synchronizedHahn echo sequence was used with a nutation frequency of 120 kHz anda spinning frequency of 60 kHz. The relaxation delay was optimizedfor each sample with optimal values in the range of 8–12 s.Spectra were referenced to adamantane set at 1.8 ppm. Spectral analysiswas performed using the DMFIT software.38

13C ssNMR spectra of 13C-enriched cathodeswere acquired on a Bruker 300 MHz Avance I spectrometer using a 2.5mm double resonance probe. A rotor synchronized Hahn echo sequencewas used with a nutation frequency of 100 kHz, a spinning frequencyof 25 kHz, and a relaxation delay of 20 s. Spectra were referencedto the adamantane tertiary group set at 38.5 ppm. Two dimensional(2D) hom*onuclear correlation experiments were acquired on a Bruker400 MHz Avance I spectrometer using a 2.5 mm probe spinning with aspinning frequency of 10 kHz. Radio Frequency Driven Recoupling (RFDR)39 was used to recouple the dipolar hom*onuclearinteractions using a nutation frequency of 105 kHz for a durationof 200 rotor cycles (20 ms).

1H–6Li heteronuclear correlation experimentswere performed on a Bruker 700 MHz Avance III spectrometer using a1.3 mm double resonance probe and a spinning frequency of 60 kHz.Cross-polarization (CP) was used to correlate 1H–6Li dipolar coupled nuclei with a 5 ms contact time.

17O spectra of discharged cathodes were acquired ona Bruker 850 and 700 MHz Avance III spectrometers using a 1.3 mm probeand a spinning frequency of 60 kHz. A rotor synchronized Hahn echosequence was used with the highest radio frequency (RF) nutation frequencyachieved equal to 91 and 96 kHz on the 700 and 850 MHz spectrometers,respectively). A relaxation delay of 1–1.5 s was used, andthe experiment times varied between 20 and 48 h on the 850 MHz spectrometer.The spectra of LiOH and Li2O were acquired on a Bruker700 MHz Avance III with a single pulse excitation (nutation frequencyof 132 kHz) and a relaxation delay of 5 s. The spectrum of Li2CO3 was acquired on a Bruker 850 MHz Avance IIIspectrometer using a 4 mm double resonance probe using a single pulseexcitation (nutation frequency of 42 kHz) with a relaxation delayof 15 s. Natural abundance 17O spectra of anhydrous lithiumacetate (CH3O2Li) and lithium formate (HCO2Li) were acquired on a Bruker 900 MHz Avance II spectrometer,using a 4 mm double resonance probe with a rotor synchronized Hahnecho (nutation frequency 42 kHz). A natural abundance 17O spectrum of Li2O2 was acquired on a Bruker850 MHz Avance III spectrometer using a static probe with a solidecho pulse excitation (62 kHz nutation frequency; 30 s relaxationdelay). MAS spinning frequencies are specified in the figure captions.Spectra were fit using either SPINEVOLUTION40 spin dynamics simulation program or the line shape analysis toolin the Bruker software Topspin.

3. Resultsand Discussion

a. 17O NMR SpectralLibrary of PossibleReaction Products

In order to identify the electrochemicalproducts formed in the cathode, we first obtain 17O spectra(Figure (Figure1a–f)1a–f) and relevant NMR parameters(from fits to the spectra; Table 1) from anassembled library of possible reaction products. For lithium peroxideand carbonate, the fits to the 17O spectra were supportedby density functional theory calculations of the NMR parameters, asdescribed in ref (33). As the natural abundance of the 17O isotope is only0.034%, isotope enrichment was used where possible, and high fieldmeasurements were employed for increased sensitivity.

Monitoringthe Electrochemical Processes in the Lithium–AirBattery by Solid State NMR Spectroscopy (3)

17O experimental NMR spectraof the various model compounds(black) and simulated best fits (gray) (b–f); the simulatedspectrum shown in (a) was calculated using the NMR parameters obtainedfrom 17O-enriched Li2O2 obtainedelectrochemically:33 (a) Static spectrumof natural abundance Li2O2, (b) MAS (12.5 kHz)spectrum of 10% 17O enriched Li2CO3, (c) MAS (12.5 kHz) of natural abundance CH3CO2Li, (d) MAS (60 kHz) spectrum of 10% 17O enriched LiOH,(e) MAS (12.5 kHz) of natural abundance HCO2Li, and (f)MAS (60 kHz) spectrum of 10% 17O enriched Li2O.

The naturalabundance spectrum of lithium peroxide (Figure (Figure1a)1a) was acquired without MAS due to the large Cq value expected from ab initio calculations of the NMR parameters.33 While the low signal-to-noise (S/N) ratio doesnot allow accurate values of the isotropic chemical shift and quadrupolarcoupling to be extracted, the width of this spectrum of more than2000 ppm in a 20 T field is consistent with the quadrupole singularitiescalculated with NMR parameters determined from the electrochemicallyformed lithium peroxide, as shown in our previous work.33 The large coupling constant (18 MHz), whichis more than twice that of most 17O species measured inthe solid state,41 makes it easy to distinguishLi2O2 from the other possible products. Therelatively sharp resonance between 0 and 200 ppm seen in the naturalabundance Li2O2 spectrum is due to a lithiumcarbonate impurity. The two crystallographically distinct oxygen sitesof lithium carbonate, the most common electrolyte decomposition product,are well resolved in its 17O MAS spectrum (Figure (Figure1b),1b), and can be fit with second order quadrupoleline shapes with Cq values of the orderof 7.2 and 7.4 MHz. The Cq of anotherpossible decomposition product LiOH (Figure (Figure1d)1d) is similar (7.05 MHz) but its chemical shift is very different,−15 ppm vs 154 and 174 ppm for the carbonate (note that theobserved center of masses of these resonances are shifted to higherfrequencies from their isotropic chemical shift values due to thesecond order quadrupolar shift). The natural abundance spectra oflithium acetate and formate (Figure (Figure1c,e)1c,e) donot have sufficient S/N, even at the high field strength of 21.2 T,to allow accurate determination of their quadrupolar coupling constants.Nevertheless, a Cq value of the orderof 7.5 MHz can be used to fit the line width with a chemical shiftof the order of 265 ppm, both values lying within the respective rangesreported for carboxyl functional groups.41 Finally, lithium oxide has been suggested as a discharge productformed in TEGDME electrolytes based on XRD data.42 No quadrupolar broadening is expected due to the cubicenvironment of the oxygen sites, and a narrow and well-resolved 17O resonance of Li2O at 35.6 ppm (Figure (Figure1f)1f) is observed. Thus, if present, its 17O signature can be easily used to identify it in the discharge products.Products containing ether functional groups were not measured, buttheir 17O resonances are expected to have isotropic shiftsin the range of 0–100 ppm with Cq values of the order of 11 MHz.41

Table 1

Experimental and Calculated 17O NMR Parametersof the Various Discharge Products Obtained fromthe Fits to the Spectra Shown in Figure Figure11

δiso/ppm|Cq|/MHzη
compoundexperimentDFTexperimentDFTexperimentDFT
Li2O222722318.0(0.2)*18.60.00(0.04)0
Li2CO3 (O1)1741597.20(0.05)7.40.94(0.05)0.97
Li2CO3 (O2)1541397.40(0.05)7.70.88(0.05)0.91
LiOH–157.05(0.05)0.10(0.1)
Li2O350
HCO2Li∼2657.3(1.6)
CH3CO2Li∼2657.7(1.5)

δiso is the isotropicchemicalshift, Cq is the quadrupole coupling constant,and η is the quadrupole asymmetry parameter. Values in bracketscorrespond to the error in determining the EFG parameters. *The Li2O2 value was extracted from a static spectrum ofLi2O2 obtained electrochemically.33

As the measurements of possible electrochemical productswere performedat several different field strengths and with various MAS frequencies,it is not straightforward to compare them with spectra obtained ofproducts formed in an operating battery. Therefore we have simulatedthe spectra using the parameters listed in Table 1, for all compounds at a single field, 20 T, and spinningfrequency, 60 kHz (Figure (Figure2).2). The simulatedspectra demonstrate that high-field 17O MAS spectroscopyis an effective tool for identifying and distinguishing the variousoxygen functionalities. Separating between lithium carboxylate groupssuch as formate and acetate, for example, will, however, require additional 1H and/or 13C measurements.

Monitoringthe Electrochemical Processes in the Lithium–AirBattery by Solid State NMR Spectroscopy (4)

Simulationof the 17O central transition line shapeof the various products at 20 T and 60 kHz MAS with the NMR parametersgiven in Table 1.

b. Electrochemistry

In order to identifythe products formed in working lithium–oxygen batteries, cellswere assembled as described in the Experimental Section. The discharge process was limited by either the discharge capacity(1000 mAh/g; referred to as partial discharge) or by a voltage limitationof 2 V allowing the cell to fully discharge. Representative electrochemicalplots for cells discharged to 1000 mAh/g and with a discharge voltagelimit of 2 V are shown in Figure Figure3.3. For characterizationby 17O NMR, the cells were cycled under an 17O enriched oxygen atmosphere and spectra were obtained followingpartial and full discharge and on charging to 4.5 V after dischargeto a voltage of 2 V. For 1H studies, additional stateswere investigated as indicated in Figure Figure33.

Monitoringthe Electrochemical Processes in the Lithium–AirBattery by Solid State NMR Spectroscopy (5)

Electrochemical profilesof two representative Li–oxygencells, one that was limited to a discharge capacity of 1000 mAh/g(dashed gray) and a second cell that was fully discharged to 2 V (black). 17O NMR spectra were acquired for cathodes at the state ofcharge indicated by black circles and 1H NMR spectra wereacquired for those indicated by both black and gray circles.

c. 17O NMR of Products Formed in CycledCathodes

The 17O spectra of the partially andfully discharged and partially charged cathodes are shown in Figure Figure4a–c4a–c along with simulations (fits) of theline shapes using the parameters determined for the possible productsformed in the cathode (from Table 1). The maindischarge product seen on full discharge (corresponding to a capacityof 2000 mAh/g) is lithium peroxide (Figure (Figure4b),4b), corresponding to approximately 80% of the spectral intensity,on the basis of the spectral simulation. However, the spectra wereacquired with a relatively short relaxation delay that is shorterthan the longitudinal relaxation time, T1, of the various possible expected products, and thus the actualamount of lithium peroxide is expected to deviate slightly from thisvalue. A detailed analysis of this effect as well as a discussionof the effect of the RF excitation bandwidth and variation in Cq on the relative phase fractions is providedin the Supporting Information. Taking the T1 relaxation into consideration, the relativeamount of lithium peroxide is estimated as 83% with an error of 17%(mainly due to the error in determining the T1 constant).In addition to the peroxide, a measurable amount of lithium carbonate(5.0 (0.4)%) of the total signal intensity (number in brackets representsthe error), lithium hydroxide (10 (5)%) and lithium formate (2.0 (0.2)%)are seen. The resonance at around 260 ppm was assigned to lithiumformate and not acetate based on the 1H spectra presentedin a later section. The 1H spectra also allow for moreaccurate quantification of the LiOH concentration. Similar measurementsperformed at a lower field of 16.4 T could be fit with similar ratiosof products (within a few percent) supporting the fits of the 20 Tspectra. All the products detected in this 17O measurementmust be a result of a reaction between one of the battery components(lithium ions, cathode and electrolyte) with the 17O enrichedoxygen gas or one of its 17O-enriched products. It is worthnoting that the lithium carbonate and hydroxide resonances do notshow the well-defined second order quadrupole line shapes expectedfrom crystalline solids. In the case of the carbonate, similar 17O line shapes were observed in spectra of crystalline carbonateat elevated temperatures due to rotation of the CO32– on the microsecond time scale,43 suggesting that the carbonate ions may be mobile, and possiblyindicating the formation of a disordered/amorphous carbonate phase.

Monitoringthe Electrochemical Processes in the Lithium–AirBattery by Solid State NMR Spectroscopy (6)

17O Hahn echo (20 T) spectra of the products formedin cathodes at different state of charge: discharge to 1000 mAh/g(a), discharge to 2 V (b), and discharge to 2 V (capacity of 1650mAh/g) followed by charge to 4.5 V (capacity of 1070 mAh/g) (c). Theexperimental spectra (black lines) were simulated (dashed gray) byusing the relevant Li2O2 (blue), Li2CO3 (pink), LiOH (green), and HCO2Li (red)NMR parameters. The spectra have been scaled so that their heightsare equal.

The spectrum acquired from a partiallydischarged cathode (limitedto a discharge capacity of 1000 mAh/g) reveals a similar distributionof electrochemical products to that detected at full discharge withLi2O2 (80 (17)%), Li2CO3 (6.0 (0.5)%), LiOH (12 (6)%), and LiO2CH (2.0 (0.2)%).Only small differences are seen in the relative concentrations ofspecies, which suggests that these four products are formed simultaneouslythroughout the discharge process rather than forming at differenttimes during the discharge. Thus, despite the improved cycling performanceachieved when limiting the depth of discharge,4,44 asimilar extent of electrolyte decomposition is observed (relativeto peroxide formation) for partial and full discharge.

The 17O spectrum of the fully discharged and then chargedto 4.5 V cathode (Figure (Figure4c)4c) shows that a largeamount of the peroxide has decomposed, decreasing in relative contributionto 56 (13)%, with lithium carbonate now corresponding to 16 (2)% ofthe total spectral intensity, hydroxide to 15 (8)%, and formate toabout 13 (1)%. Thus, as the voltage increases to 4.5 V, most, butnot all, of the lithium peroxide has been decomposed. (A method toallow the spectral intensity of the three different spectra to becompared is discussed later.) Previous studies have shown that carboncathodes preloaded with lithium peroxide exhibited a charge voltageof 4.15 V, leading to almost full removal of the lithium peroxide.45 Our results indicate that indeed some of theperoxide was removed at lower voltages but the formation of additionaldecomposition products must increase the overpotential of the chargeprocess beyond 4 V. This is consistent with DEMS measurements performedfollowing the discharge of lithium–oxygen cells where mostlyO2 evolution was detected on charge below 4 V vs lithium,while significant CO2 release was observed at higher chargingpotentials between 4 and 4.5 V.22,23

In summary, 17O NMR measurements reveal that the mainelectrochemical product in the first discharge is lithium peroxide,but with a non-negligible contribution from the 17O-enricheddecomposition products lithium carbonate, hydroxide, and formate and/oracetate.

d. 1H ssNMR of Cycled Cathodes

To monitor the hydrogen containingspecies, 1H Hahn-echoNMR spectra were acquired from cathodes extracted from cells cycledto various stages (Figure (Figure3).3). The spectra aredominated by a resonance at 2.6 ppm, which corresponds to the PVDFbinder in the cathode. Examining first cells that were fully discharged(Figure (Figure5a),5a), we can identify two main products:a broad peak in the range −1.0 to −1.5 ppm assignedto LiOH, and a weaker resonance at 8 ppm assigned to lithium formate.Another resonance can be resolved at about 0.5 ppm, which will bediscussed shortly. As the cells are charged, the LiOH and 0.5 ppmpeak decrease in intensity until they almost completely disappearwhen the cell is half charged (charge capacity of 1000 mAh/g). Theformate peak intensity on the other hand does not change noticeablyas the cell is charged and is still present at midcharge. In the spectrumof the cell following the second discharge (top row; D2v-C4.65-D2v)the intensity of all the signals grow, indicating that even more significantdecomposition occurs with further cell cycling.

Monitoringthe Electrochemical Processes in the Lithium–AirBattery by Solid State NMR Spectroscopy (7)

1H Hahn echo spectra of cathodes at different statesof charge following (a) full discharge to 2 V and (b) partial dischargeto 1000 mAh/g. Spectra are labeled according to the state of charge,where D and C stand for discharge and charge, respectively, and eitherthe voltage or capacity limit are specified. Spectra have been normalizedto the intensity of the PVDF resonance, which is assumed to be constant.

Spectra acquired from cells thatwere stopped at partial discharge(Figure (Figure5b)5b) display similar 1H resonanceswith lithium hydroxide resonating in the range of −1.5 to 1ppm and lithium formate at 8 ppm. The main difference observed withdepth of discharge is the disappearance of the LiOH resonance at lowercharging voltages, (4.3 V for partially discharged cathodes as comparedwith 4.5–4.6 V for cathodes that were fully discharged). Thisdifference can be ascribed to the thinner layer of insulating productsformed on the cathode surface at partial discharge, which leads toa lower overpotential on charge.

The PVDF signal, although usefulas a reference for quantifyingthe amounts of products formed, obscures any additional weaker resonanceswith chemical shifts in a similar shift range, for example, lithiumacetate (1.9 ppm), lithium methoxide (3.5 ppm), and other possibleether fragments (approximately, 3 ppm). These resonances can potentiallybe identified by subtracting the pristine cathode spectrum from thatof the cycled ones, allowing the signal at around 0.5 ppm to be resolvedmore clearly and an additional resonance at 3.5 ppm can be identifiedin the spectra of the partially discharged cells (Figure S5, Supporting Information). Further support forthe presence of these signals is provided by a 1H–6Li correlation NMR spectra acquired from a cathode dischargedto 1000 mAh/g vs 6Li metal to form 6Li-enrichedelectrochemical products (Figure (Figure6).6). This experimentallows signals from protons and 6Li nuclei in close spatialproximity to be correlated by using cross-polarization (CP)46 for magnetization transfer from 1H to 6Li, and it therefore selects products that containboth nuclei in close proximity. Although the various species cannotbe resolved in the 6Li dimension, four species can be clearlyresolved in the 1H projection: lithium formate at 8 ppm,lithium hydroxide at −1 ppm and the two additional environmentsat 3.5 ppm and 0.5 ppm. Based on solution NMR measurements of cathodeswashed with D2O (Figure S6, SupportingInformation), we assign the 3.5 ppm peak to the dilithium saltformed from the central fragment of DME, LiOCH2CH2OLi. This signal is observed in significant amounts only in cathodesthat were partially discharged to 1000 mAh/g and for cathodes thatwere only partially charged after this partial discharge. It disappearson charging to 4.3 V (Figure S5, Supporting Information). Its absence from the 1D difference and 2D heteronuclear correlationspectra of cathodes that were fully discharged suggests that thisdilithium salt reacts further as the discharge process proceeds, possiblyoxidizing to lithium formate and lithium hydroxide (see below). The0.5 ppm environment is assigned to a disordered lithium hydroxidephase: while crystalline, stoichiometric lithium hydroxide gives riseto 1H resonance at −1.4 ppm (Figure S7a, Supporting Information); the hydroxide signalsdetected from the cathode span −1 to 1 ppm. This shift in resonancefrequency is tentatively assigned to the disorder in the hydroxidephase. To confirm this assignment, we monitored the 7Lisignal build up in a cross-polarization experiment as a function ofthe cross-polarization time from surrounding protons (Figure S7c). The similar time scale of the buildup rate of the signals at −1 and 0.5 ppm confirms they belongto phases that are structurally similar at least on a short length-scale.

Monitoringthe Electrochemical Processes in the Lithium–AirBattery by Solid State NMR Spectroscopy (8)

1H–6Li 2D heteronuclear correlationof a cathode discharged to 1000 mAh/g. The spectrum was acquired at16.4 T (1H Larmor frequency 700 MHz) at 60 kHz MAS usinga 5 ms cross-polarization time to transfer magnetization from 1H (vertical) to 6Li (horizontal) spectra nuclei.

The 1H ssNMR measurements are consistent with the 17O spectra, showing the formation of lithium hydroxide uponcathode discharge and its full removal during the charging process.In addition, lithium formate, which has been detected in former studies,31 is shown here to accumulate in the cathode uponcycling. The 1,2-ethanediol lithium salt is observed at partial discharge,but it then decomposes on charging the cell to 4.3 V. It has almostcompletely disappeared in a cathode allowed to fully discharge.

e. 13C ssNMR of 13C EnrichedCathodes

In order to identify the sources of lithium carbonateformation, cells were cycled with cathodes that were prepared with 13C enriched amorphous carbon. While these cathodes may differin their performance from that of superP Li carbon (due to differencesin properties such as the nature of the surface groups, particle sizesand surface area), they are used here to investigate the carbon stabilityin the relevant electrochemical window. A representative electrochemicalprofile is shown in Figure S8 in the SupportingInformation.

The 13C MAS spectra (Figure (Figure7)7) all contain the signal of the bulk electrode at130 ppm, a typical shift position for an sp2 hybridizedcarbon. As the cell is fully discharged, a new weak carbon resonanceappears at 168 ppm corresponding to lithium carbonate. Upon charge,the carbonate signal grows noticeably, but is completely removed atfull charge when a voltage of just over 4.5 V is reached. Significantlylarger quantities of carbonate are formed on the second dischargeand accumulate on the cathode by the end of two cycles. On the fifthdischarge, the amount of carbonate formed corresponds to about 9%of the total carbon signal detected.

Monitoringthe Electrochemical Processes in the Lithium–AirBattery by Solid State NMR Spectroscopy (9)

13C Hahn echo MAS NMR spectraacquired from cycled 13C enriched cathodes. The positionof the carbonate groupof Li2CO3 is indicated via gray dashes, alongwith its integrated intensity relative to the total carbon signal.

Since a lithium carbonate signalis not detected when using thenatural abundance (nonenriched) carbon electrode, for similar experimentalparameters (and direct excitation of the 13C resonances),the carbonate signal detected in these measurements must be enrichedin 13C and thus originate from the decomposition of thecarbon electrode rather than the aprotic solvent. This signal shouldbe contrasted with the carbonate signal detected in the 17O spectra that can originate from the electrolyte or carbon cathodeor both. One possibility, supported by prior DEMS, FTIR, and carbonatedissolution reactions also using 13C enriched electrodes,16,24 is that the carbonate is formed from the reaction between lithiumperoxide and the carbon electrode, especially at the higher voltagesreached during charge.

To investigate the location of the carbonatespecies, a 13C hom*onuclear correlation experiment was performed(Figure (Figure8),8), which allows us to detect spatialproximity between 13C sites. A hom*onuclear correlationwas detected between thelithium carbonate resonance at 168 ppm, and the carbon electrode signalat 130 ppm, which shows that at least some of the lithium carbonateforms directly (within a few angstroms) on the surface of the carbonelectrode and not on top of the layer of peroxide in contact withthe electrolyte. This result is consistent with the observed 13C enriched carbonate being a product of the reaction betweenthe electrode and lithium peroxide. Interestingly, the 13C signal of the carbonate signal nearby the carbon electrode thatgives rise to the cross-peaks is at a slightly higher frequency thanthe sharp carbonate signal. The shift is tentatively ascribed to electroniceffects due to the interaction with the carbon electrode and/or thedisorder of the carbonate at this interface.

Monitoringthe Electrochemical Processes in the Lithium–AirBattery by Solid State NMR Spectroscopy (10)

13C–13C 2D hom*onuclear correlationof a 13C carbon enriched cathode on its 5th discharge to2 V. The spectrum was acquired at 9.4 T (Larmor frequency 100 MHz),with 10 kHz MAS, and an RFDR mixing time of 20 ms.

Measurements were also performed using DME 20% 13C-enrichedon the methyl group. No 13C signal could be detected fromfully discharged cathodes in experiments using direct excitation of 13C resonances, and 1H to 13C and 7Li to 13C CP experiments, indicating that the electrolytedecomposition products formed from the CH3 group of DME(formate and carbonate) are present in too low concentrations to bedetected over the natural abundance carbon signal. A weak signal fromthe lithium formate could be detected in a 1H to 13C CP spectrum at the end of the second discharge, consistent withthe 1H NMR experiments, which show that this species accumulateson cycling. Further experiments are in progress using DME with higher 13C enrichment levels to obtain spectra with better signal-to-noise.

f. Summary of Electrochemical Reactions in theFirst Cycle in DME

We can now combine all of the informationcollected from various NMR measurements described above and estimatethe extent of formation and decomposition of the various detectedproducts (Figure (Figure9).9). The quantities of lithiumhydroxide, formate, and 1,2-ethanediol lithium salt are based on the 1H spectra where the PVDF signal was used as internal reference.The relative lithium peroxide and carbonate amounts were extractedfrom the 17O data (correcting for the relaxation effectas discussed in the Supporting Information) and were determined by comparing their relative signal intensityto that of the 17O signals of the protonated species lithiumformate. Due to the relatively large error in determining the amountof lithium hydroxide from the 17O NMR data, only 1H spectra were used to quantify its formation.

Monitoringthe Electrochemical Processes in the Lithium–AirBattery by Solid State NMR Spectroscopy (11)

Summary of the products formed in thefirst cycle of the cell inmoles per gram of carbon electrode: Li2O2 (blue),Li2CO3 (pink), LiOH (green), HCO2Li (red), and (LiOCH2)2 (purple). Data collectedfrom cathodes with a discharge limit of 2 V is given by filled circlesand a limit of 1000 mAh/g in open circles. The whole range is shownin (a) and an enlargement of the gray area is plotted in (c). Representativevoltage profiles are shown in (b).

At partial discharge(limited to 1000 mAh/g), five products can be identified: lithiumperoxide as the major product, and four decomposition products, LiOH,lithium formate, lithium carbonate and the 1,2-ethanediol lithiumsalt. The decomposition pathway of ether solvents has been discussedin several publications.10,4749 While it is mostly agreed that lithium peroxide is formed on dischargeby a two-step process,50

Monitoringthe Electrochemical Processes in the Lithium–AirBattery by Solid State NMR Spectroscopy (12)

i

Monitoringthe Electrochemical Processes in the Lithium–AirBattery by Solid State NMR Spectroscopy (13)

ii

several possibilities have been raisedforthe mechanism for the decomposition of the ether electrolyte, whichdiffer in the nature of the attacking group. Among the suggestionsare reactions involving the superoxide ion and molecular oxygen,10,51 autoxidation by molecular oxygen,52 andreactions with the peroxide product.48,53 DEMS and coulometrymeasurements have shown that the electron to oxygen ratio (e/O2) consumed in the discharge reaction in DME based electrolytesis always larger than two.24 Since twoelectrons are required for the formation of lithium peroxide, it wasconcluded that any parasitic reactions involving the electrolyte areprobably due to the reactivity of the strong nucleophile Li2O2 (and not the intermediate LiO2).24 DFT calculations performed by Bryantsev et al.have demonstrated that it is unlikely that DME decomposition is initiatedby nucleophilic attack of the superoxide ion, these calculations beingsupported by reactivity tests of DME with KO2.54 DME decomposition pathways by reaction withsolid lithium peroxide are supported by the high reactivity of theperoxide surface groups.55 However, severalmechanistic pathways can occur, resulting in different decompositionproducts. Recent computational studies have shown that hydrogen abstractionfrom the methylene group of DME is energetically favorable, assumingreactivity of single oxygen bridging sites on the surface of the peroxidespecies.49 An experimental study basedon in situ electrochemical quartz crystal microbalance (EQCM), solutionphase NMR, and matrix-assisted laser desorption/ionization time-of-flight(MALDI-TOF) measurements ascribed the reactivity in triglyme electrolytesto the peroxide anion (LiO2) speciesformed prior to precipitation of the solid Li2O2.48

Some plausible reactions for the formation of the observedproductsin the current study (with 17O enriched O2)are

Monitoringthe Electrochemical Processes in the Lithium–AirBattery by Solid State NMR Spectroscopy (14)

iii

Monitoringthe Electrochemical Processes in the Lithium–AirBattery by Solid State NMR Spectroscopy (15)

iv

Monitoringthe Electrochemical Processes in the Lithium–AirBattery by Solid State NMR Spectroscopy (16)

v

Monitoringthe Electrochemical Processes in the Lithium–AirBattery by Solid State NMR Spectroscopy (17)

vi

Monitoringthe Electrochemical Processes in the Lithium–AirBattery by Solid State NMR Spectroscopy (18)

vii

Monitoringthe Electrochemical Processes in the Lithium–AirBattery by Solid State NMR Spectroscopy (19)

viii

The formation of the diglyme lithiumsalt,LiOCH2CH2OLi, at partial discharge (eq iii), which is only detected by 1H NMRand not by 17O NMR spectroscopy, supports the mechanismssuggested by Aurbach et al., involving a reaction between the peroxideand the DME α-carbon (reaction path b, Scheme 2 in ref (48)). Presumably, the methylperoxylithium product formed in reaction iii, CH3OOLi, is not detected in our study, because it is readilyoxidized to lithium formate (reaction iv) viaformaldehyde as an intermediate. Reaction iv also forms water, which can react further to form LiOH (reaction v). Note that the decomposition of one molecule ofDME results, via these reactions, in two molecules of water and fourmolecules of LiOH, helping to explain why LiOH is a dominant producton discharge. On deeper discharge, the diglyme lithium salt has alower concentration, presumably due to its further reaction to formother lithium salts such as carbonate (reaction vi), hydroxide and formate. Formate is present throughout (in low concentrations)as it is both a product and a reactant in the decomposition pathwayssuggested above. The reaction of the carbon electrode to form lithiumcarbonate (reaction viii) is not significanton the first discharge, as very little carbonate is seen in the 13C spectra of 13C enriched electrodes at this stage,consistent with the DEMS and FTIR study of Thotiyl et al.16 Upon charge, the lithium hydroxide graduallydecomposes and is completely removed when the voltage increases upto 4.5 V. Lithium peroxide is decomposed, partially in a reversiblemanner to release O2 and to a lesser extent by reactingwith the carbon electrode to form lithium carbonate (as detected by 13C NMR):

Monitoringthe Electrochemical Processes in the Lithium–AirBattery by Solid State NMR Spectroscopy (20)

ix

This reaction produces only two electronsfrom three peroxide molecules (as compared to six electrons for fulloxidation of the peroxide). While the lithium peroxide accounts forabout 80% of the discharge capacity (1600 mAh/g), on the basis ofthe total capacity and the Li2O2

:decompositionproducts ratio, its signal intensity observed in the 17O spectrum quickly drops by about 80% when the cell is charged from2 to 4.5 V with a charge capacity of only about 500 mAh/g. Oxidationof the carbon electrode can at least partially explain this fast decreasein peroxide intensity, the decomposition of Li2CO3 at 4.5 V, and above releasing the residual four electrons (fromthe original three peroxide anions):

Monitoringthe Electrochemical Processes in the Lithium–AirBattery by Solid State NMR Spectroscopy (21)

x

Although essentially allthe carbonate decomposesat higher voltages in the first cycle, in subsequent cycles, thisis not complete (as observed by 13C NMR). Interestingly,the 2D

13C NMR spectrum of the 13C enricheddischarged electrode after five cycles shows two types of carbonatesignals: one nearby the carbon, and a sharper resonance that is furtherfrom the carbon (i.e., it does not give rise to a carbon–carbonatecross peak) and is less disordered. It is likely that it is this lattercarbon that is more difficult to decompose electrochemically on chargeand gradually accumulates on cycling. Finally, in contrast to thecarbonate, lithium formate, which is mainly formed upon discharge,maintains an essentially constant level and is not removed upon charge.The accumulation of the formate and at a later stage lithium carbonatemay be one cause of the capacity fading observed in operating cells.

4. Conclusions

The electrochemical productsformed upon cycling of lithium–oxygencells in 1 M LiTFSI in DME have been identified and monitored usingsolid state NMR. While lithium peroxide is identified as the maindischarge product, non-negligible electrolyte decomposition is detectedalready at capacities of 1000 mAh/g. The formation of peroxide anddecomposition products continues until the end of discharge whereadditional carbonate formation is detected from the reaction betweenperoxide and the carbon electrode. 13C hom*onuclear correlationexperiments reveal the formation of carbonate directly on the electrodesurface and provide evidence for the carbon electrode reactivity.Upon charge, significant decomposition of lithium peroxide occursat voltages below 4 V to form oxygen and via a significant electrochemicalreaction with the electrode to form lithium carbonate. Lithium hydroxideis removed at voltages, lower than 4.5 V, while for complete removalof the carbonate voltages higher than 4.5 V are required. Lithiumformate does not completely decompose under the conditions used hereand it accumulates on the cathode. Upon further cycling, additionalelectrolyte decomposition occurs, resulting in the accumulation oflithium salts on the carbon surface leading to capacity fading andincreased overpotentials.

Solid state NMR, in particular 17O NMR, is demonstratedto be a valuable tool in the assessment of lithium–air cellsand is expected to be useful in the evaluation of the functionalityof new cell components such as electrolytes, electrode materials,and catalytic species.

Acknowledgments

The UK 850 MHzSolid-State NMR Facility used in this researchwas funded by EPSRC and BBSRC, as well as by the University of Warwickwith partial funding through Birmingham Science City Advanced MaterialsProjects 1 and 2 supported by Advantage West Midlands (AWM) and theEuropean Regional Development Fund (ERDF). We also thank the NationalUltrahigh-Field NMR Facility for Solids (Ottawa, Canada), a nationalresearch facility funded by the Canadian Foundation for Innovation,the Ontario Innovation Trust, and Recherche Quebec, for access tothe 900 MHz NMR spectrometer. We thank Prof. Dominic Wright and MatthewDunstan for their help in synthesizing the 17O enrichedcompounds and Prof. Peter Bruce and Timothy King for helpful discussions.M.L is an awardee of the Weizmann Institute of Science – NationalPostdoctoral Award for Advancing Women in Science and thanks the EUMarie Curie intra-European fellowship for funding. We thank JohnsonMatthey for funding (A.J.M.) and fruitful discussions. This researchwas supported by the Engineering and Physical Sciences Research Councilas part of the Supergen energy storage consortium.

Supporting Information Available

Further details on the analysisof the 17O NMR data; additional experimental results supporting 1H spectra assignment; electrochemical performance of 13C enriched electrodes. This material is available free ofcharge via the Internet at http://pubs.acs.org.

Notes

The authors declare nocompeting financial interest.

Supplementary Material

References

  • Bruce P. G.; Freunberger S. A.; Hardwick L. J.; Tarascon J. M.Li-O2 and Li-S Batterieswith High Energy Storage. Nat. Mater.2012, 11, 19–29. [PubMed] [Google Scholar]
  • Ogasawara T.; Débart A.; Holzapfel M.; Novák P.; Bruce P. G.Rechargeable LI2O2 Electrodefor Lithium Batteries. J. Am. Chem. Soc.2006, 128, 1390–3. [PubMed] [Google Scholar]
  • Girishkumar G.; McCloskey B.; Luntz a. C.; Swanson S.; Wilcke W.Lithium–AirBattery: Promise and Challenges. J. Phys. Chem.Lett.2010, 1, 2193–2203. [Google Scholar]
  • Christensen J.; Albertus P.; Sanchez-Carrera R. S.; Lohmann T.; Kozinsky B.; Liedtke R.; Ahmed J.; Kojic A.A Critical Review ofLi/Air Batteries. J. Electrochem. Soc.2012, 159, R1. [Google Scholar]
  • Garcia-Araez N.; Novák P.Critical Aspects in the Developmentof Lithium–airBatteries. J. Solid State Electrochem.2013, 17, 1793–1807. [Google Scholar]
  • Mizunu F.; Nakanishi S.; Kotani Y.; Yokoishi S.; Iba H.RechargeableLi–Air Batteries with Carbonate-Based Liquid Electrolytes. Electrochemistry2010, No. 5, 78. [Google Scholar]
  • Freunberger S. a; Chen Y.; Peng Z.; Griffin J. M.; Hardwick L. J.; Bardé F.; Novák P.; Bruce P. G.Reactions in theRechargeable Lithium–O2 Battery with Alkyl CarbonateElectrolytes. J. Am. Chem. Soc.2011, 133, 8040–7. [PubMed] [Google Scholar]
  • McCloskey B. D.; Bethune D. S.; Shelby R. M.; Girishkumar G.; Luntz A. C.Solvents’ Critical Role inNonaqueous Lithium–OxygenBattery Electrochemistry. J. Phys. Chem. Lett.2011, 2, 1161–1166. [PubMed] [Google Scholar]
  • Xu W.; Xu K.; Viswanathan V. V; Towne S. A.; Hardy J. S.; Xiao J.; Nie Z.; Hu D.; Wang D.; Zhang J.Reaction Mechanisms for the LimitedReversibility of Li–O2 Chemistry in Organic CarbonateElectrolytes. J. Power Sources2011, 196, 9631–9639. [Google Scholar]
  • Freunberger S. a; Chen Y.; Drewett N. E.; Hardwick L. J.; Bardé F.; Bruce P. G.The Lithium-OxygenBattery with Ether-Based Electrolytes. Angew.Chem., Int. Ed. Engl.2011, 1–6. [PubMed] [Google Scholar]
  • Lu Y.; Kwabi D. G.; Yao K. P. C.; Harding J. R.; Zhou J.; Zuin L.; Shao-Horn Y.The DischargeRate Capability ofRechargeable Li–O2 Batteries. Energy Environ. Sci.2011, 4, 2999. [Google Scholar]
  • Lu Y.-C.; Gallant B. M.; Kwabi D. G.; Harding J. R.; Mitchell R. R.; Whittingham M. S.; Shao-Horn Y.Lithium–Oxygen Batteries:Bridging Mechanistic Understanding and Battery Performance. Energy Environ. Sci.2013, 6, 750. [Google Scholar]
  • Jung H.; Hassoun J.; Park J.; Sun Y.; Scrosati B.An ImprovedHigh-Performance Lithium–Air Battery. Nat. Chem.2012, 4, 579–85. [PubMed] [Google Scholar]
  • Read J.; Mutolo K.; Ervin M.; Behl W.; Wolfenstine J.; Driedger a.; Foster D.Oxygen TransportProperties of OrganicElectrolytes and Performance of Lithium/Oxygen Battery. J. Electrochem. Soc.2003, 150, A1351. [Google Scholar]
  • Xu D.; Wang Z.; Xu J.; Zhang L.; Zhang X.Novel DMSO-BasedElectrolyte for High Performance Rechargeable Li-O2 Batteries. Chem. Commun. (Cambridge, U. K.)2012, 48, 6948–50. [PubMed] [Google Scholar]
  • OttakamThotiyl M. M.; Freunberger S. a; Peng Z.; Bruce P. G.The CarbonElectrode in Nonaqueous Li-O2 Cells. J. Am. Chem. Soc.2013, 135, 494–500. [PubMed] [Google Scholar]
  • Chen Y.; Freunberger S. a; Peng Z.; Bardé F.; Bruce P. G.Li–O2 Battery with a DimethylformamideElectrolyte. J. Am. Chem. Soc.2012, 134, 7952–7. [PubMed] [Google Scholar]
  • Veith G. M.; Dudney N. J.Current Collectorsfor Rechargeable Li–Air Batteries. J.Electrochem. Soc.2011, 158, A658. [Google Scholar]
  • Nasybulin E.; Xu W.; Engelhard M. H.; Nie Z.; Burton S. D.; Cosimbescu L.; Gross M. E.; Zhang J.Effects ofElectrolyte Salts on thePerformance of Li–O2 Batteries. J. Phys. Chem. C2013, 117, 2635–2645. [Google Scholar]
  • Du P.; Lu J.; Lau K. C.; Luo X.; Bareño J.; Zhang X.; Ren Y.; Zhang Z.; Curtiss L. a; Sun Y.-K.; et al. Compatibility of Lithium Salts with Solvent of theNon-Aqueous Electrolyte in Li–O2 Batteries. Phys. Chem. Chem. Phys.2013, 15, 5572–81. [PubMed] [Google Scholar]
  • Younesi R.; Hahlin M.; Treskow M.; Scheers J.; Johansson P.; Edström K.Ether Based Electrolyte, LiB(CN)4 Salt andBinder Degradation in the Li–O2 Battery Studiedby Hard X-ray Photoelectron Spectroscopy (HAXPES). J. Phys. Chem. C2012, 116, 18597–18604. [Google Scholar]
  • Gallant B. M.; Mitchell R. R.; Kwabi D. G.; Zhou J.; Zuin L.; Thompson C. V.; Shao-Horn Y.Chemical and Morphological Changesof Li–O2 Battery Electrodes Upon Cycling. J. Phys. Chem. C2012, 116, 20800–20805. [Google Scholar]
  • McCloskey B. D.; Speidel A.; Scheffler R.; Miller D. C.; Viswanathan V.; Hummelshøj J. S.; Nørskov J. K.; Luntz A. C.Twin Problems ofInterfacial Carbonate Formation in Nonaqueous Li–O2 Batteries. J. Phys. Chem. Lett.2012, 3, 997–1001. [PubMed] [Google Scholar]
  • McCloskey B. D.; Bethune D. S.; Shelby R. M.; Mori T.; Scheffler R.; Speidel A.; Sherwood M.; Luntz A. C.Limitations in Rechargeabilityof Li–O2 Batteries and Possible Origins. J. Phys. Chem. Lett.2012, 3, 3043–3047. [PubMed] [Google Scholar]
  • Mitchell R. R.; Gallant B. M.; Shao-Horn Y.; Thompson C. V.Mechanisms of MorphologicalEvolution of Li2O2 Particles During ElectrochemicalGrowth. J. Phys. Chem. Lett.2013, 4, 1060–1064. [PubMed] [Google Scholar]
  • Débart A.; Bao J.; Armstrong G.; Bruce P. G.An O2 Cathode for RechargeableLithium Batteries: The Effect of a Catalyst. J. Power Sources2007, 174, 1177–1182. [Google Scholar]
  • Lu Y.-C.; Gasteiger H. a.; Parent M. C.; Chiloyan V.; Shao-Horn Y.The Influenceof Catalysts on Discharge and Charge Voltages of Rechargeable Li–OxygenBatteries. Electrochem. Solid-State Lett.2010, 13, A69. [Google Scholar]
  • Shao Y.; Park S.; Xiao J.; Zhang J.; Wang Y.; Liu J.Electrocatalysts for Nonaqueous Lithium–AirBatteries: Status,Challenges, and Perspective. ACS Catal.2012, 2, 844–857. [Google Scholar]
  • Oh S. H.; Nazar L. F.Oxide Catalystsfor Rechargeable High-Capacity Li–O2 Batteries. Adv. Energy Mater.2012, 2, 903–910. [Google Scholar]
  • McCloskey B. D.; Scheffler R.; Speidel A.; Bethune D. S.; Shelby R. M.; Luntz A. C.On the Efficacy of Electrocatalysis in Nonaqueous Li–O2 Batteries. J. Am. Chem. Soc.2011, 133, 18038–41. [PubMed] [Google Scholar]
  • Freunberger S. a; Chen Y.; Drewett N. E.; Hardwick L. J.; Bardé F.; Bruce P. G.The Lithium–OxygenBattery with Ether-BasedElectrolytes. Angew. Chem., Int. Ed. Engl.2011, 50, 8609–13. [PubMed] [Google Scholar]
  • Lim H.; Yilmaz E.; Byon H. R.Real-TimeXRD Studies of Li–O2 Electrochemical Reaction inNonaqueous Lithium–OxygenBattery. J. Phys. Chem. Lett.2012, 3, 3210–3215. [PubMed] [Google Scholar]
  • Leskes M.; Drewett N. E.; Hardwick L. J.; Bruce P. G.; Goward G. R.; Grey C. P.Direct Detectionof Discharge Products in Lithium–OxygenBatteries by Solid-State NMR Spectroscopy. Angew.Chem., Int. Ed. Engl.2012, 51, 8560–3. [PubMed] [Google Scholar]
  • Xu Z.; Stebbins J. F.6Li Nuclear Magnetic Resonance ChemicalShifts, Coordination Number and Relaxation in Crystalline and GlassySilicates. Solid State Nucl. Magn. Reson.1995, 5, 103–12. [PubMed] [Google Scholar]
  • Xiao J.; Hu J.; Wang D.; Hu D.; Xu W.; Graff G. L.; Nie Z.; Liu J.; Zhang J.-G.Investigation of the Rechargeabilityof Li–O2 Batteries in Non-Aqueous Electrolyte. J. Power Sources2011, 196, 5674–5678. [Google Scholar]
  • Huff L. A.; Rapp J. L.; Zhu L.; Gewirth A. A.Identifying Lithium–AirBattery Discharge Products through 6Li Solid-State MASand 1H–13C Solution NMR Spectroscopy. J. Power Sources2013, 235, 87–94. [Google Scholar]
  • Abys J. A.; Barnes D. M.; Feller S.; Rouse G. B.; Risen W. M. Jr.Preparation of O17-Labelled Glasses andGlass Precursers. Mater. Res. Bull.1980, 15, 1581–1587. [Google Scholar]
  • Massiot D.; Fayon F.; Capron M.; King I.; Le Calv S.; Alonso B.; Durand J.-O.; Bujoli B.; Gan Z.; Hoatson G.Modelling One- andTwo-Dimensional Solid-State NMRSpectra. Magn. Reson. Chem.2002, 40, 70–76. [Google Scholar]
  • Bennett A. E.; Griffin R. G.; Ok J. H.; Vega S.Chemical Shift CorrelationSpectroscopy in Rotating Solids: Radio Frequency-Driven Dipolar Recouplingand Longitudinal Exchange. J. Chem. Phys.1992, 96, 8624. [Google Scholar]
  • Veshtort M.; Griffin R. G.SPINEVOLUTION: A Powerful Tool forthe Simulation ofSolid and Liquid State NMR Experiments. J. Magn.Reson.2006, 178, 248–282. [PubMed] [Google Scholar]
  • Wu G.Solid-State 17O NMR Studiesof Organic and Biological Molecules. Prog. Nucl.Magn. Reson. Spectrosc.2008, 52, 118–169. [PubMed] [Google Scholar]
  • Laoire C. O.; Mukerjee S.; Plichta E. J.; Hendrickson M. a.; Abraham K. M.Rechargeable Lithium/TEGDME-LiPF6/O2 Battery. J. Electrochem. Soc.2011, 158, A302. [Google Scholar]
  • Dunstan M. T.; Griffin J. M. ; Blanc F.; Leskes M.; Grey C. P.. Lithium Ion Dynamics in Li2CO3 Studied by Solid State NMR and First PrinciplesCalculations. In Preparation.
  • Oh S. H.; Black R.; Pomerantseva E.; Lee J.; Nazar L. F.Synthesisof a Metallic Mesoporous Pyrochlore as a Catalyst for Lithium–O2 Batteries. Nat. Chem.2012, 4, 1004–10. [PubMed] [Google Scholar]
  • Meini S.; Tsiouvaras N.; Schwenke K. U.; Piana M.; Beyer H.; Lange L.; Gasteiger H. A.Rechargeability of Li–AirCathodes Pre-filled with Discharge Products Using an Ether-Based ElectrolyteSolution: Implications for Cycle-Life of Li–Air Cells. Phys. Chem. Chem. Phys.2013, 11478–11493. [PubMed] [Google Scholar]
  • Stejskal E..; Schaefer J.; Waugh J. .Magic-Angle Spinning and PolarizationTransfer in Proton-Enhanced NMR. J. Magn. Reson.1977, 28, 105–112. [Google Scholar]
  • Wang H.; Xie K.Investigation of OxygenReduction Chemistry in Ether and CarbonateBased Electrolytes for Li–O2 Batteries. Electrochim. Acta2012, 64, 29–34. [Google Scholar]
  • Sharon D.; Etacheri V.; Garsuch A.; Afri M.; Frimer A. a.; Aurbach D.On the Challenge of Electrolyte Solutions for Li–AirBatteries: Monitoring Oxygen Reduction and Related Reactions in PolyetherSolutions by Spectroscopy and EQCM. J. Phys.Chem. Lett.2013, 4, 127–131. [PubMed] [Google Scholar]
  • Assary R.S.; Lau K. C.; Amine K.; Sun Y.-K.; Curtiss L. A.Interactionsof Dimethoxy Ethane with Li2O2 Clusters andLikely Decomposition Mechanisms for Li–O2 Batteries. J. Phys. Chem. C2013, 117, 8041–8049. [Google Scholar]
  • Hardwick L. J.; Bruce P. G.The Pursuit of Rechargeable Non-Aqueous Lithium–OxygenBattery Cathodes. Curr. Opin. Solid State Mater.Sci.2012, 16, 178–185. [Google Scholar]
  • Bryantsev V. S.; Uddin J.; Giordani V.; Walker W.; Addison D.; Chase G. V.The Identificationof Stable Solvents for NonaqueousRechargeable Li–Air Batteries. J. Electrochem.Soc.2012, 160, A160–A171. [Google Scholar]
  • Bryantsev V. S.; fa*glioni F.PredictingAutoxidation Stability of Ether- and Amide-BasedElectrolyte Solvents for Li–Air Batteries. J. Phys. Chem. A2012, 116, 7128–38. [PubMed] [Google Scholar]
  • Assary R. S.; Lau K. C.; Amine K.; Sun Y.-K.; Curtiss L. a.Interactionsof Dimethoxy Ethane with Li2O 2 Clusters andLikely Decomposition Mechanisms for Li–O2 Batteries. J. Phys. Chem. C2013, 117, 8041–8049. [Google Scholar]
  • Bryantsev V. S.; Blanco M.Computational Study of the Mechanisms of Superoxide-InducedDecomposition of Organic Carbonate-Based Electrolytes. J. Phys. Chem. Lett.2011, 2, 379–383. [Google Scholar]
  • Lau K. C.; Assary R. S.; Redfern P.; Greeley J.; Curtiss L. A.ElectronicStructure of Lithium Peroxide Clusters and Relevance to Lithium–AirBatteries. J. Phys. Chem. C2012, 116, 23890–23896. [Google Scholar]

Articles from The Journal of Physical Chemistry. C, Nanomaterials and Interfaces are provided here courtesy of American Chemical Society

Monitoring
the Electrochemical Processes in the Lithium–Air
Battery by Solid State NMR Spectroscopy (2024)
Top Articles
Latest Posts
Article information

Author: Aracelis Kilback

Last Updated:

Views: 6658

Rating: 4.3 / 5 (44 voted)

Reviews: 83% of readers found this page helpful

Author information

Name: Aracelis Kilback

Birthday: 1994-11-22

Address: Apt. 895 30151 Green Plain, Lake Mariela, RI 98141

Phone: +5992291857476

Job: Legal Officer

Hobby: LARPing, role-playing games, Slacklining, Reading, Inline skating, Brazilian jiu-jitsu, Dance

Introduction: My name is Aracelis Kilback, I am a nice, gentle, agreeable, joyous, attractive, combative, gifted person who loves writing and wants to share my knowledge and understanding with you.