Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active...

43
www.sciencemag.org/content/350/6260/530/suppl/DC1 Supplementary Materials for Cycling Li-O 2 batteries via LiOH formation and decomposition Tao Liu, Michal Leskes, Wanjing Yu, Amy J. Moore, Lina Zhou, Paul M. Bayley, Gunwoo Kim, Clare P. Grey* *Corresponding author. E-mail: [email protected] Published 30 October 2015, Science 350, 530 (2015) DOI: 10.1126/science.aac7730 This PDF file includes: Materials and Methods Supplementary Text Figs. S1 to S23 References (43–47)

Transcript of Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active...

Page 1: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

www.sciencemag.org/content/350/6260/530/suppl/DC1

Supplementary Materials for

Cycling Li-O2 batteries via LiOH formation and decomposition Tao Liu, Michal Leskes, Wanjing Yu, Amy J. Moore, Lina Zhou, Paul M. Bayley,

Gunwoo Kim, Clare P. Grey*

*Corresponding author. E-mail: [email protected]

Published 30 October 2015, Science 350, 530 (2015) DOI: 10.1126/science.aac7730

This PDF file includes:

Materials and Methods Supplementary Text Figs. S1 to S23 References (43–47)

Page 2: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

*

This file includes:

Materials and Methods

Additional Experiments and Discussion

1. SEM and Electrochemical Investigation of the Three Different Electrodes (Fig. S1 and S2)

2. Comparison of the Capacity Obtained for the I-/I3

- couple (in Ar) with that Obtained for the

Li-O2 Cells (Fig. S3)

3. 7Li NMR Characterization of the Discharge Products in the Presence of LiI and O2 (Fig. S4)

4. 7Li,

1H

NMR and SEM Characterization of the Discharge Products in the Absence of LiI

(Fig.S5)

5. SEM Characterization of the Discharge Products Obtained with LiI (Fig. S6)

6. SEM and ssNMR Characterization of the Products with LiI in TEGDME (Fig. S7)

7. SEM Characterization of Electrodes after Multiple Cycles (Fig. S8)

8. Li-O2 Battery Using an SP Carbon Electrode and LiI/DME (Fig. S9)

9. Electrochemistry of Samples for XRD and NMR in Fig. 2 (Fig. S10)

10. Li-O2 Batteries Cycled at Higher Rates (Fig. S11)

11. Li-O2 Cells Cycled in the Presence of High Concentrations of Water (Fig. S12)

12. Added Water Leads to the Formation of even Larger LiOH Crystals (Fig. S13)

13. Further Comments on the Specific Energy Density Calculations

14. Establishing the Discharge Mechanisms

14.1. Sources of H in the Formed LiOH and the Role of Water (Fig. S14-18)

14.2. LiOH/electron Molar Ratio during Discharge (Fig. S19)

15. Establishing the Charge Mechanisms (Fig. S20-21)

16. Electrochemistry Data and Fitted NMR Spectra for Figure S19 (Fig. S22-23)

Page 3: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

2

Materials and Methods:

Materials

1,2-Dimethoxyethane (DME) (Sigma Aldrich, 99.5%) and tetraethylene glycol dimethyl ether

(TEGDME) (Sigma Aldrich, 99%) solvents were was refluxed with calcium hydride and distilled

under a nitrogen atmosphere, and then stored over 4 Å molecular sieves under argon. The final

water content of the solvents was measured to be below 10 ppm by Karl Fischer titration

(Metrohm 899). Deuterated DME (Sigma Aldrich, D-10, 99.5 D%, 99% CP) was used

as-received. Molecular sieves were washed with ethanol and acetone, dried overnight in an oven

at 70°C and then at 275°C in vacuo for two days. Lithium bis(trifluoromethyl)sulfonylimide

(LiTFSI) (3M FluoradTM

, HQ115) and LiI (Sigma-Aldrich, 99.9%) were dried at 160°C and

200°C, respectively, in vacuo for 12 hours before being used to prepare the electrolyte. Super P

(SP) carbon black (~50 nm) and TiC nanoparticles (~40 nm) were purchased from Timcal and

Skyspring nanomaterials respectively. All materials were stored and handled in an Ar glovebox

with <0.1 ppm O2 and <0.1 ppm H2O.

Methods

Electrode fabrication methods

Mesoporous SP carbon electrodes were prepared from a mixture of 24 wt% SP carbon black,

38 wt% polyvinylidene fluoride (PVDF, copolymer) binder, and 38 wt% dibutylphthalate (DBP,

Sigma-Aldrich) in acetone. The slurry was then spread into a self-supporting film and cut into ½

inch in diameter discs, which were washed with diethyl ether to remove the DBP. The resulting

films were then annealed at 120°C in vacuo for 12 hours and transferred to the glovebox without

exposure to air. The final carbon content in the electrodes is 39 wt%. Mesoporous TiC electrodes

were prepared via a similar procedure, with carbon being replaced by TiC. A mixture of 48 wt%

TiC, 8 wt% PVDF and 44 wt% DBP was used with acetone to make the slurry. The subsequent

film making and drying method are exactly the same as used for fabricating SP carbon

electrodes.

Aqueous graphene oxide solution was synthesized by a modified Hummer’s method (43).

Briefly, concentrated H2SO4 (96 ml) was added to a mixture of graphite flakes (2 g) and sodium

nitrate (2 g), which was stirred at 0°C in a water/ice bath. KMnO4 (12 g) was then gradually

Page 4: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

3

added and the mixture was continuously stirred at 0°C for 90 minutes. The reaction temperature

was subsequently raised and kept at 35°C for 2 hours, after which 80 ml deionized water was

slowly added to the suspension. Additional water (200 ml) and H2O2 (30%, 10 ml) were

introduced. At this point, a suspension of graphite oxides was obtained. This graphite oxides

suspension was allowed to settle down and the clear solution at the top was repeatedly removed

and replaced with deionized water until the suspension became neutral. The resulting slurry of

graphite oxide was subjected to many ultrasonication and centrifugation cycles until no sediment

was found at the bottom of the centrifuge tube. A well-dispersed aqueous graphene oxide

solution was finally synthesized.

To fabricate reduced graphene oxide (rGO) electrodes, the obtained graphene oxide solution

was first concentrated by annealing it in a vial at 80-100°C to form a viscous gel that has a

graphene oxide concentration of ~10 mg/ml. The gel was cast onto a stainless steel (ss) mesh

(Advent Research Materials) using a volumetric pipette and then frozen and stored in a vial in

liquid N2. The graphene oxide electrodes on ss-meshes were freeze-dried for 12 hours in vacuo

and then subjected to pyrolysis in a furnace under Ar at 550°C for 2 hours, to obtain rGO

electrodes. These electrodes were further dried at 150°C in vacuo before being used to make

batteries. The masses of rGO electrodes were carefully measured by comparing the masses of a

few bare meshes with their respective masses after the rGO electrodes had been coated, an

average value being taken for a specific batch of electrodes.

Li-O2 cell assembly and electrochemical measurements

All Li-O2 cells investigated in this work are based on a Swagelok design. They were

assembled by stacking a disc of lithium foil (0.38 mm thickness, Sigma-Aldrich), 2 pieces of

borosilicate glass fibre separators (Whatman) soaked with electrolyte, and an O2 positive

electrode (SP, TiC or rGO). The electrolytes used in this study include 0.25 M LiTFSI/DME or

0.25 M LiTFSI/TEGDME with/without the addition of 0.05 M LiI. A 0.5 cm2 hole was drilled

through the current collector, so that the positive electrode can readily access O2. The assembled

Swagelok cell was then placed in a 150 ml Li-O2 glass chamber, the two electrodes being

electrically connected to two tungsten feedthroughs. Pure O2 was purged through the chamber

via two Young’s taps for 25 minutes, and the cell was then rested for 10 hours before cycling.

The volume of TEGDME electrolyte used for a cell was typically 0.2 ml. DME electrolyte is

Page 5: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

4

more volatile and was found to evaporate rapidly during the O2 purge and to be absorbed into the

viton rubber coating of the electrical cables (this latter problem occurs due to the current

in-house design of our Li-O2 cells and can be avoided using better cell designs). Consequently,

0.7-1 ml DME electrolyte was used for a cell. The electrode loading in this work ranged from

0.01 to 0.15 mg and the thickness of the rGO electrodes varied from 30 to 200 μm. For Li-O2

cells subjected to more prolonged cycling, thinner electrodes (30-50 μm) were used. The cycling

rate was quoted based on the mass of carbon in an electrode. For example, 5 A/gc rate (Fig. 4B)

of a 0.01 mg electrode is equivalent to cycling the cell at a current of 50 uA, giving a rate per

unit area of 0.1 mA/cm2. The electrochemical measurements (Galvanostatic discharge/charge,

cyclic voltammetry) were conducted using either an Arbin battery cycler or a Biologic VMP. All

potentials are referenced against Li/Li+.

Electrode characterization

For all batteries, the characterization of electrodes involved first disassembling the cell,

rinsing the O2 positive electrode twice in dry acetonitrile (<1 ppm H2O), each time with 2 ml

CH3CN for 30 minutes. The washed electrodes were then dried in vacuo overnight for further

characterization.

X-ray diffraction (XRD) measurements were performed using a Panalytical Empyrean

diffractometer operated in a reflection mode, with Cu Kα1 radiation (λ=1.5406 Å). The cycled

electrode was sandwiched between two Kapton polyimide films in an air tight sample holder.

Scanning electron microscopic (SEM) images were recorded with a Hitachi S-5500 in lens

field emission electron microscope. The electrode samples were hermetically sealed during

transfer to the electron microscope. Once the seal was opened, the sample was loaded into the

high vacuum SEM chamber within 10 seconds.

All solid-state NMR (ssNMR) spectra were acquired on either a 16.4 T Bruker Avance III or

an 11.7 T Bruker Avance III spectrometer using 1.3 mm HX probes. A rotor synchronized

Hahn-echo pulse sequence was used to acquire 1H magic angle spinning (MAS) spectra with a

spinning speed of 60 kHz (unless stated otherwise), and an rf field strength of 125 kHz. A

one-pulse sequence was used to acquire 7Li NMR spectra under MAS and static conditions, with

an rf field strength of 167 kHz. 1H and

7Li shifts were externally referenced to solid adamantane

at 1.8 ppm and lithium carbonate at 0 ppm, respectively. The same receiver gain, number of

Page 6: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

5

scans and recycle delay values (optimized values of between 10-20 s were used) were employed

to measure the electrodes (from the same batch) to allow quantitative comparison between

spectra.

Additional Experiments and Discussion

1. SEM and Electrochemical Investigation of the Three Different Electrodes

In this work, the electrochemical performance of cells made of the three types of electrodes

discussed on the main text were compared to evaluate the magnitude of various contributions to

the charge overpotential, such as catalytic effects of different electrode materials, concentration

polarization due to the diffusion of electrode active species and ohmic loss caused by the

insulating discharge product. The rGO electrodes, figure S1 (c), clearly contain much larger pore

sizes and pore volumes than SP electrodes (e), which leads to a lower tortuosity and thus more

efficient diffusion of the active species within the electrolyte (Li+, solvated O2, mediators etc.) in

rGO than in SP. Therefore the smaller overpotential for rGO electrode than SP is ascribed, at

least in part, to the interconnecting macroporous framework. SP carbon (e) and TiC (f) electrodes

are comprised of particles of similar sizes (~50 nm) and are made by the same fabrication

method. They both have similar mesoporous electrode structures. The difference in the

electrochemical performance between SP and TiC electrodes (Fig. 1A, main text) is tentatively

attributed to the difference in their intrinsic catalytic activities.

It is important to stress that it is difficult to separate unambiguously the different contributions to

the overpotential (e.g., the activation barrier of the reaction, ohmic loss, diffusion of active

species). When comparing the SP with TiC electrodes, for example, it is difficult to ensure that

the electrical resistance and surface areas of the electrodes are identical even if the pore structure

is similar. Thus, the current method of comparison (Fig. 1) gives a qualitative estimation of the

various origins for the charge overpotentials rather than a quantitative evaluation.

Page 7: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

6

Fig. S1 SEM images showing hierarchically macroporous rGO electrodes at various magnifications (a-d),

mesoporous Super P (SP) carbon (e) and mesoporous TiC electrodes (f).

Page 8: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

7

As can be seen on in figure S2 (a), rGO and SP carbon electrodes exhibit good stability within

the voltage window 2.4-3.5 V; gradually rising cathodic and anodic currents were observed out

of this voltage range. TiC is a less inert electrode material in LiTFSI/DME: rapidly rising

cathodic and anodic currents were observed below 2.5 and 3.75 V, respectively. Figures S2 (b)

and (c) illustrate that rGO, SP and TiC electrode all reversibly cycle LiI (3I- ↔ I3

- + 2e

-). In

figure S2 (b), the separation between the redox peaks of I-/ I3

- in TEGDME-based electrolyte

(blue curve) is wider than that in DME-based electrolyte (red curve). This is probably associated

with the higher viscosity of TEGDME and hence a slower diffusion of mediators in this

electrolyte.

Fig. S2 (a) Cyclic voltammograms of cells using rGO, Super P (SP) and TiC electrodes in 0.25 M

LiTFSI/DME under an Ar atmosphere; (b) cyclic voltammograms comparing cells using rGO electrodes

in 0.05 M LiI/0.25 M LiTFSI/DME and TEGDME electrolytes under an Ar atmosphere; (c) cyclic

voltammograms of cells using SP and TiC electrodes in 0.05 M LiI/0.25 M LiTFSI/DME under an Ar

atmosphere. Sweep rate for all cells was 5 mV/s.

Page 9: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

8

2. Comparison of the Capacity Obtained with the I-/I3

- Couple (in Ar) with that Obtained

for the Li-O2 Cells

The capacity of a Li-iodine redox battery is typically evaluated based on the mass of iodine (the

active material), which gives a theoretical capacity of 211 mAh/g (44) i.e., [(96485/3.6) mA]/127

g. In our cells with the TEGDME electrolyte, the number of moles of I- was 2.1×10

-5 mol, i.e.,

2.7×10-3

g. The electrical charge extracted from the SP, TiC and rGO cells under an Ar

atmosphere was 2.5×10-4

, 5×10-4

and 7×10-4

mAh, respectively, giving only 0.09, 0.18, 0.26

mAh/gI, which is much lower than the theoretical capacity based on the I-/I3

- couple and the total

I- present in the cell. This indicates that the majority of the iodide ions did not participate in the

electrochemical reaction. This is not, however, surprising, as no effective convection is available

in our cells. As a result, the capacity is solely dependent upon the self-diffusion of electroactive

species. Similar values are obtained for cells using the DME electrolyte, being 0.18, 0.18, 0.07

mAh/gI for SP, TiC and rGO cells, respectively.

In a Li-O2 cell investigated in this work, the capacity is calculated based only on the mass of

electrode material (SP carbon, TiC, or rGO). To compare the capacities obtained with and

without O2, we also calculated the capacity of LiI cells based on the mass of the electrode

materials, as illustrated in figure S3 below. It is clear that the specific capacities of all 3

electrodes cycled under Ar are much smaller than those of the Li-O2 batteries, when using the

same electrode materials.

Page 10: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

9

Fig. S3 Galvanostatic charge-discharge curves of cells cycled with 0.05 M LiI in 0.25 M LiTFSI /

TEGDME and DME electrolytes, in an Ar rather than O2 atmosphere; these cells were first charged and

then discharged. The grey line in each graph shows the corresponding electrodes discharged in the same

DME-based electrolyte in an O2 atmosphere.

Page 11: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

10

3. 7Li NMR Characterization of the Discharge Products in the Presence of LiI and O2

Fig. S4 Comparison between the 7Li static NMR spectra, acquired at 11.7 T, of a discharged rGO

electrode from a Li-O2 cell using a 0.05 M LiI/0.25 M LiTFSI/DME electrolyte and those of the model

compounds LiOH, Li2CO3 and Li2O2.

The characteristic quadrupolar line shape of the discharged rGO electrode sample (Fig. S4)

overlaps with that of LiOH, rather than Li2CO3 or Li2O2, suggesting the discharge product is

overwhelmingly LiOH.

Page 12: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

11

4. 7Li,

1H

NMR and SEM Characterization of the Discharge Products in the Absence of

LiI

As can be seen in figure S5 (a), the rGO electrode without the added LiI (blue curve) exhibits a

lower discharge capacity than that with added LiI (red curve). The higher capacity results from

the much larger concentration of discharge products (LiOH) that more efficiently take up the

pore volume in macroporous rGO electrodes in the latter case (see Fig. S6). The black curve in

(a) represents a cell with an rGO electrode in 0.05 M LiI/0.25 M LiTFSI/DME electrolyte

galvanostatically discharged in an Ar atmosphere: its capacity is negligible compared to that

cycled in an O2 atmosphere (red curve).

In figure S5(b), a single resonance at 0 ppm in the 7Li MAS ssNMR spectrum (acquired at 16.4

T) and the absence of satellite transition peaks in the 7Li static ssNMR spectrum suggest that

Li2O2 is the predominant discharge product (Fig. S4) when LiI is absent. 1H MAS ssNMR

measurement shows a resonance at -1.5 ppm, suggesting LiOH is also present in the discharge

products. The resonances at 2.3 and 8 ppm are attributed to residual DME solvent and lithium

formate, respectively, in the electrode. SEM images (c-d) of this discharged electrode reveal that

the rGO electrode surfaces are fully covered by toroidal particles (~500 nm), the toroids being a

characteristic morphology for Li2O2, consistent with the ssNMR measurements.

Page 13: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

12

Fig. S5 The electrochemistry of a Li-O2 battery using a rGO electrode in a 0.25 M LiTFSI/DME

electrolyte (a, blue curve) and characterization of this discharged rGO electrode by ssNMR (b) and SEM

(c-d). The electrochemistry of a Li-O2 battery discharged in the presence of LiI is shown for comparison

in (a) (red curve). This battery was interrupted at a capacity of 3.2 mAh. The black curve in (a) shows the

LiI battery discharged under Ar; its capacity is negligible.

Page 14: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

13

5. SEM Characterization of the Discharge Products obtained with LiI

Fig. S6 SEM images of a fully discharged rGO electrode in 0.05 M LiI/0.25 M LiTFSI/DME electrolyte.

Numerous large particles fill up the pores of the rGO electrode (a). When we cut the electrode and

investigated the interior space (b-d), flower-like particles larger than 15 μm were also observed. Some of

these particles grow on the insulating glass fiber separators (d), indicating that these LiOH crystals were

formed via a solution precipitation process.

Page 15: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

14

6. SEM and ssNMR Characterization of the Products with LiI in TEGDME

When discharged at a slow rate of 100 mA/gc, a voltage plateau at 2.7 V was observed for a

Li-O2 cell made with TEGDME-based electrolyte (Fig. S7 (a)). In the ssNMR spectra (b) of the

discharged rGO electrode, acquired at 16.4 T, resonances at -1.5 and 8.3 ppm in the 1H MAS

spectrum indicate the existence of lithium hydroxide and lithium formate, respectively, lithium

hydroxide being the dominant product; the resonance at 4.8 ppm is attributed to water and those

at 3.3 and 0.7 ppm are ascribed to residual TEGDME solvent in the electrode. The single

resonance at 1.0 ppm and the characteristic line shape of the quadrupolar 7Li static ssNMR

spectrum with its satellite transition peaks further confirm that LiOH is the predominant

discharge product. SEM images (c) of this discharged electrode reveal that instead of forming

large flower-like particles as seen with the DME-based electrolyte, LiOH exists as a thin film

covering the rGO electrode surface, resulting in a lower discharge capacity.

Page 16: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

15

Fig. S7 The electrochemistry (a) of a Li-O2 battery using an rGO electrode in a 0.05 M LiI/0.25 M

LiTFSI/TEGDME electrolyte and characterization of the discharged rGO electrode by ssNMR (b) and

SEM (c).

Page 17: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

16

7. SEM Characterization of Electrodes after Multiple Cycles

Fig. S8 SEM images of rGO electrodes cycled in 0.05 M LiI/0.25 M LiTFSI/DME electrolyte at the end

of the 1st, 15

th and 100

th discharge (D) and charge (C). Only trace amounts of residual LiOH are seen after

a full charge on the rGO electrode surfaces. Note that there is always some variation in the porous

structure between and in different parts of the electrodes, even from the same batch, because of the

casting method. The SEM images therefore vary to a degree between samples and across the sample due

to these changes in porosity. For example, the image of the 100th cycle charged sample shows a very

porous part of the electrode, whereas 1st and 15

th cycle charge electrodes represent less porous ones. We

note that the porous elongated features (arrows) on the 100th cycle charged electrode belong to the porous

graphene electrode, not to the remaining LiOH.

Page 18: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

17

8. Li-O2 Battery Using an SP Carbon Electrode and LiI/DME

When discharged at 70 mA/gc, a voltage plateau at 2.65 V (Fig. S9(a)) was observed for a Li-O2

cell made using an SP carbon electrode and LiI-added DME electrolyte. In the ssNMR spectra

(b), the dominant resonance at -1.5 ppm in 1H and a single resonance at 1.0 ppm in the

7Li

ssNMR spectra suggest that LiOH is the main discharge product; this is further corroborated by

the 7Li static spectrum. The resonance at 4.8 ppm in the

1H ssNMR spectrum is attributed to

water and those at 3.3, 2.6 and 1.0 ppm are ascribed to residual DME solvent in the electrode and

saturated carbons. SEM images (c) of this discharged electrode reveal that LiOH exhibit disc and

sheet-like morphologies. Notably, these discs/sheets of LiOH are ~500 nm in size, much smaller

than those observed in rGO electrodes (Fig. S6) even though the same electrolyte was used.

After full charge, much of the surface of the SP electrode became bare again, although some

regions containing residual LiOH were still observed. This observation suggests that LiOH can

indeed be removed during charge in the SP electrode/electrolyte system, even if it is not

complete.

Page 19: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

18

Fig. S9 The electrochemistry (a) of a Li-O2 battery using an SP carbon electrode in a 0.05 M LiI/0.25 M

LiTFSI/DME electrolyte, and characterization of the discharged SP electrode by ssNMR spectroscopy and

SEM (c), where spectra were acquired at 11.7 T (b).

Page 20: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

19

9. Electrochemistry of samples for XRD and NMR Shown in Fig. 2

Fig. S10 The electrochemistry of the two samples used for XRD and NMR measurements in fig. 2 of the

main text. A slow rate of 7 μA was used for both rGO cells.

Page 21: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

20

10. Li-O2 Batteries Cycled at Higher Rates

When cycling the cell at the higher rate of 8 A/gc the cell voltage gradually polarizes with cycle

number, probably due to more side reactions occurring at the more reducing (discharge) and

oxidizing (charge) electrochemical potentials, and the incomplete removal of discharge product.

As a result, the rGO electrode surface (c) remains covered after just 40 cycles by particles with

different morphologies from those seen for LiOH formed at lower cycling rates.

Fig. S11 The first 42 discharge-charge curves of a Li-O2 battery cycled at 8 A/gc rate using an rGO

electrode in 0.05 M LiI/0.25 M LiTFSI/DME electrolyte (a) and the corresponding terminal voltages as a

functional of cycle number (b). An SEM image (c) of the rGO electrode from the Li-O2 cell in (a) after 42

cycles. (d) represents a Li-O2 cell that was cycled for 1000 cycles with a capacity limited to 1000 mAh/gc

and then deliberately subjected to 15 much deeper discharge-charge cycles with a reversible capacity of

22,000 mAh/gc at a 1 A/gc rate. Apart from the larger voltage polarization due to the prior cycling, the cell

still demonstrates a good reversibility.

Page 22: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

21

11. Li-O2 Cells Cycled in the Presence of High Concentrations of Water

Fig. S12 The discharge-charge profiles of Li-O2 batteries cycled using rGO electrodes in a 0.05 M

LiI/0.25 M LiTFSI/DME electrolyte: (a) with 45,000 ppm of water deliberately added to the electrolyte;

(b) purged with wet O2 gas that had been passed through a water bubbler.

Experiments were performed to explore the sensitivity of this cell to water. No noticeable

difference was seen in the electrochemical performance, compared to cells cycled with no

additional water added, which demonstrates that the cell is insensitive to H2O contamination at

least at the levels investigated here (37 mg H2O per 783 mg of DME, i.e., 45,000 ppm H2O).

Page 23: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

22

12. Added Water Leads to the Formation of even Larger LiOH Crystals

Fig. S13 SEM images of an rGO electrode from a Li-O2 cell, prepared using a 0.05 LiI/0.25 M LiTFSI/DME

electrolyte with added water (20,000 ppm, 15 mg). Porous LiOH crystals of more than 30 μm were observed.

Nuclei of secondary LiOH crystals were observed to grow on the sheets of large LiOH crystals, which also

support a solution mechanism for the formation of LiOH in the system, because the large LiOH crystals are

insulating.

Page 24: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

23

13. Further Comments on the Specific Energy Density Calculations

The energy density quoted in the text (2.7 V × 3.2 mAh )÷1.5 mg = 5760 Wh/kg only takes the

rGO electrode and discharge products into account; the weight of other cell parts are not

considered. Compared with a commercial lithium iron phosphate (45) electrode (3.2 V, 0.13

mAh/1.5 mg), the current rGO electrode is able to provide 20 times more specific energy. In

comparison with the targeted specific energy (500 Wh/kg) of an aqueous Li-O2 battery (46), the

value achieved with the aprotic Li-O2 cell in this current work is still 10 times higher. However,

it should be noted that the 1.5 mg value (for the weight of the electrode) only contains electrode

pieces that can be removed from the stainless steel mesh. Errors in the weight of the electrode

can occur because some of the LiOH is formed on the separator, stuck in the stainless steel mesh

and/or washed away during rinsing. Thus 1.5 mg represents an underestimated weight for the

discharged electrode that corresponds to the 3.2 mAh capacity, leading in practice to slightly

lower specific energies than the quoted number, 5760 Wh/kg.

The electrode area used in this work is ~0.5 cm2, and hence 3.2 mAh corresponds to 6.4

mAh/cm2. This capacity based on an electrode of ~200 μm in thickness. The theoretical

discharge capacity will be larger than this: the LiOH crystals at the farther end of the electrode

away from the O2 reservoir were observed to be smaller in size and less dense than those at the

electrode regions directly facing O2, i.e. the reaction zone problem still exists to a degree in the

current cell configuration; hence there is still room for further improvement in the capacity of the

cell.

Although several critical issues with Li-O2 batteries have been addressed (low discharge capacity,

a large hysteresis, considerable side reaction with electrolyte, sensitivity to water), there are still

problems associated with the current battery, such as high volatility of DME electrolyte, intrinsic

dendrite growth and safety issues with the use of Li metal anode, low rate performance, and the

reaction zone problem (40). In particular, the rate performance and the reaction zone problems

are closely related, limiting the scale-up of any Li-O2 battery and motivating future work.

Page 25: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

24

14. Establishing Discharge Mechanisms

14.1. Sources of H in the Formed LiOH and the Role of H2O

The H source for LiOH is clearly a very important factor that should affect the performance of

the battery. First we evaluate if the H in the LiOH discharge product is from the surface

functional groups on rGO electrodes. The weight of the cell electrode (pristine weight=0.1 mg)

after discharge (3.2 mAh) is ~1.6 mg. XRD and NMR measurements show that the increased

weight are due to formation of LiOH crystals, i.e., 1.5 mg LiOH. To produce 1.5 mg LiOH,

0.0625 mg of H is needed, which is more than half the weight of the pristine rGO electrode.

Thus, GO is unlikely to be the H source for LiOH. Furthermore, 1H ssNNR of the pristine rGO

electrodes reveals a low proton content.

The water content of DME solvent was measured by a Karl Fischer apparatus to be less than 10

ppm. Approximately 1 ml electrolyte solvent was used for a battery, which leads to 1 cm3 ×

0.8683 g/cm3 × 10 ppm = 0.0086 mg H2O. This is two orders of magnitude less than needed (1.1

mg of H2O) to generate 1.5 mg of LiOH. Hence, it is unlikely that the trace amount of water in a

nominally dry electrolyte solvent is the source of H.

To test whether H is from moisture in air that gradually leaked into the sealed Li-O2 cell, we

discharged a Li-O2 battery inside an Ar glovebox (H2O<0.1 ppm). ssNMR measurements,

performed at 11.7 T, (figure S14) of this discharged rGO electrode showed that LiOH is still the

prevailing discharge product, precluding this possibility as well.

Page 26: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

25

Fig. S14 ssNMR spectra of an rGO electrode used in a Li-O2 cell discharged with 0.05 M LiI/0.25 M

LiTFSI/DME electrolyte inside an Ar glovebox (<0.1 ppm H2O). The dominant resonances at -1.5 ppm in

1H and 1.0 ppm in

7Li MAS spectra, respectively, and the characteristic line shape of the

7Li static

spectrum all suggest that LiOH is the prevailing discharge product of this cell. The other resonances

labeled in the 1H spectrum are attributed to residual DME in the electrode.

If the H were from DME molecules (and assuming that the molar ratio of consumed DME to

generated H is 1:1; note that we use this assumption simply to provide an order of magnitude

estimate of the proton content) to produce 1.5 mg LiOH, 62.5 μmol H is needed, i.e., 5.6 mg

DME (62.5 μmol ×90.12 g/mol) is required. This is equivalent to only 6.4 μl DME (5.6 mg/868.3

mg/cm3). In this work, around 1 ml DME was used to make a cell. Thus DME is one possible H

source for the production of LiOH during the first discharge process. To test this hypothesis, we

discharged a cell using an rGO electrode with a 0.05 M LiI/0.25 M LiTFSI/DME electrolyte,

prepared from as-received deuterated DME (D-10, 99.5 D%, 99% CP). 1H,

7Li ssNMR and XRD

measurements on this sample show that a crystalline lithium hydroxide phase together with a

significant amount of other discharge (decomposition) products is present in the electrode (Fig.

S15). This higher level of decomposition products is ascribed to lower purity level (99% CP) of

this deuterated DME solvent (the deuterated solvent being difficult to distil due to the low

quantity purchased (due to the cost)). The 2H ssNMR spectrum (Fig. S16) of this sample

confirms the presence of LiOD in the electrode, which is more visible in the characteristic

Page 27: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

26

spinning side band manifolds, rather than in the central band. However, LiOD is by no means the

dominant 2H signal in the discharged electrode produced in the presence of deuterated DME.

Fig. S15 1H and

7Li ssNMR and XRD of a Li-O2 cell discharged using an rGO electrode and a 0.05 M

LiI/0.25 M LiTFSI/DME electrolyte, prepared from as-received deuterated DME (D-10, 99.5 D%, 99%

CP). The NMR spectra of the discharged electrode and the reference compound are acquired at 11.7 T,

under (a, b) MAS speed of 60 kHz, and (c) static conditions. The 1H MAS spectrum shows a resonance at

-1.4 ppm that is assigned to LiOH indicating that it is present in the discharge products. The resonance at

8 ppm is attributed to lithium formate. The resonances at 1.4 and 3.9 ppm are due to saturated

hydrocarbons (1.4 ppm) and ethers (3.9 ppm). The XRD pattern (d) confirms that the major crystalline

phase in the discharged electrode is LiOH; the thin LiOH sheets observed by SEM (i.e., a short coherent

length along <001> direction) likely give rise to a weak/broad (001) reflection which is obscured by the

broad background from the Kapton film from the sample holder. The electrochemistry is given in Fig.

S22 (a, cell 1).

Page 28: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

27

Fig. S16 2H ssNMR spectra of the rGO electrode from the same cell discussed in figure S15. The spectra

were acquired at 11.7 T, with a MAS speed of 60 kHz; magnified spectra of the isotropic and 1st spinning

sideband are shown for clarity. The major discharge product observed in the 2H spectrum is not LiOD at

-1.5 ppm but instead gives rise to another resonance at 3.5 ppm due to ethers.

To investigate whether water can serve as an alternative hydrogen source, we added D2O to the

0.05 LiI/0.25 M LiTFSI/non-deuterated DME electrolyte and discharged a Li-O2 cell using an

rGO electrode. The 7Li (b-c) and

2H (d) spectra all show that deuterated lithium hydroxide is the

dominant discharge product (Fig. S17). The observation that LiOD is now the only signal visible

in the intense 2H spectrum (d) while only a weak minor

1H resonance at -1.5 ppm due to LiOH is

seen in the 1H spectrum (a) of the same sample confirms that LiOD is formed from the deuterons

of D2O in the electrolyte; this result thus clearly shows that water is the major source of

hydrogen in the reaction to form LiOH.

Page 29: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

28

Fig. S17. ssNMR spectra of an rGO electrode extracted from a Li-O2 cell, prepared using a 0.05 LiI/0.25

M LiTFSI/non-deuterated DME electrolyte but with added D2O (20,000 ppm, 15 mg). The ssNMR

spectra of LiOD reference compound is also shown for comparison: 1H (a),

7Li under MAS (b) and static

conditions (c), and 2H NMR spectra with a zoom (right) to show the isotopic region for clarity (d). The

resonances at 1.4 and 3.9 ppm in (a) are due to saturated hydrocarbons (1.4 ppm) and ethers (3.9

ppm).The electrochemistry data of the cell is given in Fig. S22(a, cell 2).

The observation that H2O supplies the H to the formed LiOH suggests that a significant amount

of H2O was introduced into the cell when deuterated DME was used (as discussed in figures S15

and 16). Experiments performed with deuterated DME dried by molecular sieves overnight still

showed that LiOH was the dominant phase and no trace of LiOD was observed (1H,

7Li and

2H

ssNMR measurements). H2O is still introduced from other sources. The capacity of the cell was

1.1 mAh. Assuming an electron/LiOH molar ratio of 1 and that all H in H2O goes to form LiOH

(see below), the amount of H2O needed was 0.37 mg, which corresponds to ~500 ppm of H2O in

the electrolyte.

Page 30: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

29

One possible source of H2O comes from the O2 purge of our cells. To test this, 1 ml of nominally

dry DME (<5 ppm H2O as determined in a Karl Fischer test) was sealed in a Li-O2 cell chamber

and purged for 5 minutes using the current O2 line. The water content of the resulting DME was

measured as 47 ppm showing that one source of H2O comes from water in the lines. Further

water likely arises from water sorbed on other cell components such as the separator.

Fig. S18 1H (a),

7Li (b) and

2H (c) ssNMR spectra of a Li-O2 cell prepared using an rGO electrode and

0.05 M LiI/0.25 M LiTFSI/deuterated DME (dried by molecular sieves overnight). ssNMR spectra of

LiOD reference compound are shown for comparison. An intense 1H resonance at -1.5 ppm (a) and a

single 7Li resonance at 1 ppm (b) suggest LiOH is the dominant phase in the discharge product. No

signature due to LiOD was observed in the 2H spectra (c) of this discharge electrode. The electrochemical

data of the cell is given in Fig. S22 (a, cell 3).

In summary, although the DME solvent in the electrolyte is a potential H source for the formed

LiOH during discharge, it does not seem to be the dominant one. Water is shown to serve as an

Page 31: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

30

alternative H source for LiOH in the current LiI mediated system. When water is intentionally

added to the DME-based electrolyte, it appears to preferentially supply H to form LiOH,

minimizing DME decomposition.

No appreciable amount of Li2CO3 or Li acetate was observed in our discharged or charged

electrodes. Only very little Li formate was observed in the 1H spectra of the discharged

electrodes. However, when Li2O2 forms as the dominant discharge product, a significant amount

of Li formate is present (13); this difference is probably related to the intrinsic reactivity of the

discharge product Li2O2 with a DME-based electrolyte, as suggested by previous studies (47).

Compared to Li2O2, the less reactive LiOH crystals appear to cause minimal chemical reactions

with the DME-based electrolyte. The reactivity of the intermediate LiO2, which will cause DME

decomposition has also been substantially reduced, which has been demonstrated by the Li-O2

cells cycled in the presence of additional water, as discussed in figures S17 and 18; this may also

be part of the reason why very little Li2CO3, Li acetate and formate were generated. LiOH does

not seem to react with the rGO electrode at the charge voltages (< 3.4 V) investigated here;

whether the side product/LiOH reacts with graphene to form Li2CO3 at higher voltages is

unclear. However, without the use of the mediator, we found that rGO oxidation is facile at

charge voltages beyond 4 V.

Finally, it should be emphasized that without the presence of LiI, DME electrolyte does not lead

to significant LiOH formation even in the presence of significant amounts of water; instead

Li2O2 is the dominant discharge product instead, large toroidal particles being observed. Of note,

this is consistent with prior work where water is shown to promote the formation of the Li2O2

toroidal particles, water contents of ≥ 500 ppm being required for toroid formation (42). The role

of water in the reaction helps explain why the kinetics (overpotential) of the 1st and 2

nd cycle are

essentially identical, but this may also be because water is not involved in the reaction that sets

the overpotential. When 0.05 M LiI is also added, no trace of Li2O2 is detected, LiOH becoming

the prevailing discharge product. Thus the LiI must play a role in promoting LiOH versus Li2O2

formation.

Page 32: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

31

14.2. Electron/LiOH Molar Ratio during Discharge

In the main text of the manuscript we proposed that the first step during discharge is a

one-electron electrochemical process, i.e., Li+ + e

- + O2 → LiO2, because the observed discharge

voltages using rGO electrodes were the same with and without LiI in the electrolyte (Fig. 1A,

main text), i.e., the first step is identical to that observed in the standard Li-air cell (2-3). We also

proposed that the subsequent conversion of LiO2 to LiOH is a chemical process that occurs via a

solution mechanism. It is clear that I- is involved in changing the equilibia so that LiOH is

formed rather than Li2O2, since Li2O2 is still the major product under wet conditions (our work

and work of reference 8).

Formally the reaction can be written:

4Li+ + 4e

- + 4O2 → 4LiO2 Electrochemical, E [1]

4LiO2 + 2H2O → 4LiOH + 3O2 Chemical, C [2]

Where the chemical reaction [2] is mediated via reactions involving I-.

Potential competitive reactions are:

2Li+ + 2e

- + 2O2 → 2LiO2 Electrochemical, E [3]

2LiO2 → Li2O2 + O2 Chemical, C [4]

Li2O2 +H2O → 2LiOH + 1/2O2 Chemical, C [5]

or

2Li+ + 2e

- + 2O2 → 2LiO2 Electrochemical, E [6]

2LiO2 → 2Li2O2 + O2 Chemical, C [7]

We investigated the electron-to-LiOH molar ratio during discharge by quantifying the number of

moles of LiOH in several cells discharged to different extents (with added LiI mediator), and

comparing their NMR spectra with those of a reference LiOH sample to determine the number of

moles of LiOH versus the moles of electrons consumed in the electrochemical discharge (Fig.

S19). On the basis of the 1H LiOH signals derived from electrodes with different capacities, the

LiOH/electron molar ratio is equal to 0.76.

Page 33: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

32

The method underestimates the LiOH content since we cannot pack all the LiOH within a

discharged cell into an NMR rotor (for example, rGO fragments are washed away during rinsing,

the rGO pieces stick in holes of the stainless steel mesh, LiOH is formed on the mesh itself and

on the glass fibre separators (as seen in the SEM), and LiOH is lost on rotor packing tools used

to pack NMR rotors etc.). We estimate we could lose as much as 20% of the sample in the

process. Given this, it appears reasonable that a LiOH:electron molar ratio close to 1 is observed

and the mechanism appears to occur via a reaction that consumes 1 Li+ per e

-. While the fit is not

perfect, our data does support a mechanism where each electron results in one LiOH and that

LiOH is the major product. We recognise that the parallel route to form Li2O2 results in a

LiOH/electron molar ratio of less than one (reactions [6] and [7]), but given that we do not detect

this product by XRD and NMR, we believe that the lower ratio originates from an inability to

measure all the LiOH. The current data clearly supports our proposal that LiOH is the major

product.

A major challenge is to identify the mechanism by which reaction [2] occurs. Given that

longevity of the DME electrolyte and the ability to cycle the battery for multiple cycles, we

speculate that I- must play a role in mopping up radicals, minimizing side reactions (noting that

some side reactions still occur, as seen in the 1H NMR above).

Page 34: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

33

Fig. S19 A quantification plot of moles of LiOH versus moles of electrons obtained by analysing 1H

NMR spectra as a function of capacity. All 1H NMR spectra are acquired at 11.7 T, with a MAS

frequency of 60 kHz, with a recycle delay of 200 s. To avoid any saturation effects due to the long

spin-lattice relaxation times in this system, a small flip-angle pulse of 15º is used for all NMR

experiments. The number of scans varies between 16 and 176. Spectral deconvolutions are carried out

with DMFIT. The electrochemistry data and fitted spectra are presented in Fig. S22 (b) and Fig. S23.

Page 35: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

34

15. Establishing the Charge Mechanism

In this work, we have demonstrated that LiOH can indeed be removed during charge at voltages

below 3.2 V using SEM, XRD and NMR (Fig. 2 in the manuscript). The fact that the cells are

able to cycle well without capacity fading also supports a reversible reaction involving LiOH.

Fig. S20 shows a Li-O2 cell cycled with a cut-off discharge voltage of 2.5 V. The end of

discharge in a Li-O2 battery is typically marked by the electrode surface being fully covered by

insulating discharge products or pores being so heavily clogged that diffusion of the electroactive

species becomes very sluggish (resulting in rapid polarization of the cell voltage). The

electrochemistry of the cell shown, which is discharged to the end, cycles with nearly an

identical electrochemical profile, demonstrating highly reversible electrochemistry (the 1st 10

cycles are shown here). The majority of the LiOH that forms during discharge must be removed

during the following charge, otherwise cell would polarize rapidly due to the blocking of its

pores and surface areas by LiOH. (This statement is supported by NMR spectra acquired at the

top of charge).

Fig. S20 The electrochemical profile of a Li-O2 cell discharged to the end. 2.5 V was set as the cut-off

voltage as side reactions become more severe below this point (see Fig. S2).

Page 36: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

35

To verify the reaction mechanism proposed in figure 4 of the main text, it is necessary to detect

whether there is any O2 evolution during charge. Since water minimizes DME decomposition

and has no appreciable effect on the electrochemical profile, we investigated a cell made using

0.05 M LiI/0.25 M LiTFSI/DME electrolyte with a 9800 ppm level of water.

This Li-O2 cell was first discharged and then connected to a mass spectrometer (MS), with the

gas atmosphere in the cell then replaced by a high purity argon gas. We subsequently measured

the cumulative gas signal after charge (Fig. S21(a)) instead of the transient response, the low

sensitivity of our set-up and the large dead-space preventing the transient measurement. As can

be clearly in figure S21(b) and its inset, a peak for both m/z 16 and 32 (black and blue traces)

was observed, confirming that the initial sharp peaks indeed represent O2. Following the sharp

O2 peak, there is another broad peak associated with m/z 32, but no corresponding m/z 16 peak

was seen (inset in Fig. S21(b)); thus this second peak is not due to O2 and is likely associated

with the ionization of DME vapor at the MS head. Due to the presence of water in the electrolyte

and its reaction with lithium metal anode, the water (m/z=18) and hydrogen (m/z=2) signals also

rise as soon as the valves were opened. This result thus clearly shows that O2 is generated in the

cell after charging at 3.2 V.

In summary, electrochemistry, NMR, XRD and SEM results all suggest that LiOH is removed

during charge in the LiI mediated Li-O2 battery. The electrochemical profile during charge and

the observation that O2 is produced during charge by mass spectrometry indicate that the first

step during charge is the oxidation of iodide anions to triiodide anions and the second step is a

redox chemical reaction, where lithium hydroxide is decomposed by triiodide anions generating

water and oxygen.

Page 37: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

36

Fig. S21 (a) The electrochemical profile of a Li-O2 battery made using an rGO electrode and 0.05 M

LiI/0.25 M LiTFSI/DME electrolyte with 9800 ppm of H2O content; the cell was discharged at a current

of 10 μA in O2 and charged at 15 μA under an Ar atmosphere. (b) Mass spectrometry measurement on the

gas atmosphere of the cell in (a) after it was charged under Ar. An initial polarization of the charge

voltage was observed for the cell in (a). This is occasionally observed for cells that have been subjected to

a very deep discharge, such as this one. (When a Li-O2 cell undergoes a deep discharge, most of the

electrode surface area is presumably covered by insulating discharge products and pores in the electrode

are clogged as well, making the diffusion of electro-active species very sluggish at the beginning,

resulting in the initial observed polarization). As charge continues, the voltage drops gradually and

eventually levels off, which is likely due to the removal of the discharge products.

Page 38: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

37

15. Electrochemistry Data and Fitted NMR Spectra for Figure S15-19

Fig. S22 (a) Electrochemistry corresponding to Li-O2 cells discussed in Fig. S15-16 (cell 1), Fig. S17 (cell

2) and Fig. S18 (cell 3); the initial dip in the voltage profile (cell 2) is likely to be caused by the formation

of solid-electrolyte-interface at the Li metal anode (due to presence of added excess water 20,000 ppm, 15

mg). (b) Electrochemistry of Li-O2 cells corresponding to the samples used to derive LiOH/electron molar

ratio in Fig. S19; the capacities of the 6 lowest capacity electrodes are noticeably lower than our usual

cells, and consistent with this more impurities were detected by NMR. We tentatively ascribe this to

lower concentrations of the graphene oxide solution used (the electrodes having poorer structural integrity)

and impurities in the DME electrolyte. Increasing the GO concentrations removed the problems in

subsequent cells from a new batch (see the 1.99 mAh electrode as an example).

Page 39: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

38

Fig. S23. Raw

1H NMR data for figure S19, used to quantify the moles of LiOH versus the moles of

electrons; an experimental spectrum, a total fit, and individual fits are shown as blue solid, red dashed,

and black lines, respectively. The intensity of the LiOH peak at approximately -1.5 ppm is used in the

calculations.

Page 40: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

REFERENCES 1. G. Girishkumar, B. McCloskey, A. C. Luntz, S. Swanson, W. Wilcke, Lithium−air battery:

Promise and challenges. J. Phys. Chem. Lett. 1, 2193–2203 (2010). doi:10.1021/jz1005384

2. P. G. Bruce, S. A. Freunberger, L. J. Hardwick, J. M. Tarascon, Li-O2 and Li-S batteries with high energy storage. Nat. Mater. 11, 19–29 (2011). Medline doi:10.1038/nmat3191

3. Y. C. Lu, B. M. Gallant, D. G. Kwabi, J. R. Harding, R. R. Mitchell, M. S. Whittingham, Y. Shao-Horn, Lithium–oxygen batteries: Bridging mechanistic understanding and battery performance. Energy Environ. Sci. 6, 750 (2013). doi:10.1039/c3ee23966g

4. R. R. Mitchell, B. M. Gallant, Y. Shao-Horn, C. V. Thompson, Mechanisms of morphological evolution of Li2O2 particles during electrochemical growth. J. Phys. Chem. Lett. 4, 1060–1064 (2013). Medline doi:10.1021/jz4003586

5. B. D. Adams, C. Radtke, R. Black, M. L. Trudeau, K. Zaghib, L. F. Nazar, Current density dependence of peroxide formation in the Li–O2 battery and its effect on charge. Energy Environ. Sci. 6, 1772 (2013). doi:10.1039/c3ee40697k

6. B. M. Gallant, D. G. Kwabi, R. R. Mitchell, J. Zhou, C. V. Thompson, Y. Shao-Horn, Influence of Li2O2 morphology on oxygen reduction and evolution kinetics in Li–O2 batteries. Energy Environ. Sci. 6, 2518 (2013). doi:10.1039/c3ee40998h

7. S. A. Freunberger, Y. Chen, Z. Peng, J. M. Griffin, L. J. Hardwick, F. Bardé, P. Novák, P. G. Bruce, Reactions in the rechargeable lithium-O2 battery with alkyl carbonate electrolytes. J. Am. Chem. Soc. 133, 8040–8047 (2011). Medline doi:10.1021/ja2021747

8. B. D. McCloskey, D. S. Bethune, R. M. Shelby, G. Girishkumar, A. C. Luntz, Solvents’ critical role in nonaqueous lithium-oxygen battery electrochemistry. J. Phys. Chem. Lett. 2, 1161–1166 (2011). Medline doi:10.1021/jz200352v

9. S. A. Freunberger, Y. Chen, N. E. Drewett, L. J. Hardwick, F. Bardé, P. G. Bruce, The lithium-oxygen battery with ether-based electrolytes. Angew. Chem. Int. Ed. 50, 8609–8613 (2011). doi:10.1002/anie.201102357

10. B. D. McCloskey, A. Speidel, R. Scheffler, D. C. Miller, V. Viswanathan, J. S. Hummelshøj, J. K. Nørskov, A. C. Luntz, Twin problems of interfacial carbonate formation in nonaqueous Li-O2 batteries. J. Phys. Chem. Lett. 3, 997–1001 (2012). Medline doi:10.1021/jz300243r

11. B. M. Gallant, R. R. Mitchell, D. G. Kwabi, J. Zhou, L. Zuin, C. V. Thompson, Y. Shao-Horn, Chemical and morphological changes of Li–O2 battery electrodes upon cycling. J. Phys. Chem. C 116, 20800–20805 (2012). doi:10.1021/jp308093b

12. M. M. Ottakam Thotiyl, S. A. Freunberger, Z. Peng, P. G. Bruce, The carbon electrode in nonaqueous Li-O2 cells. J. Am. Chem. Soc. 135, 494–500 (2013). Medline doi:10.1021/ja310258x

13. M. Leskes, A. J. Moore, G. R. Goward, C. P. Grey, Monitoring the electrochemical processes in the lithium–air battery by solid state NMR spectroscopy. J. Phys. Chem. C 117, 26929–26939 (2013). doi:10.1021/jp410429k

Page 41: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

2

14. D. Zhai, H. H. Wang, J. Yang, K. C. Lau, K. Li, K. Amine, L. A. Curtiss, Disproportionation in Li-O2 batteries based on a large surface area carbon cathode. J. Am. Chem. Soc. 135, 15364–15372 (2013). Medline doi:10.1021/ja403199d

15. E. Nasybulin, W. Xu, M. H. Engelhard, Z. Nie, S. D. Burton, L. Cosimbescu, M. E. Gross, J.-G. Zhang, Effects of electrolyte salts on the performance of Li–O2 batteries. J. Phys. Chem. C 117, 2635–2645 (2013). doi:10.1021/jp311114u

16. S. R. Gowda, A. Brunet, G. M. Wallraff, B. D. McCloskey, Implications of CO2 contamination in rechargeable nonaqueous Li-O2 batteries. J. Phys. Chem. Lett. 4, 276–279 (2013). Medline doi:10.1021/jz301902h

17. H. K. Lim, H. D. Lim, K. Y. Park, D. H. Seo, H. Gwon, J. Hong, W. A. Goddard 3rd, H. Kim, K. Kang, Toward a lithium-“air” battery: The effect of CO2 on the chemistry of a lithium-oxygen cell. J. Am. Chem. Soc. 135, 9733–9742 (2013). Medline doi:10.1021/ja4016765

18. Y. Liu, R. Wang, Y. Lyu, H. Li, L. Chen, Rechargeable Li/CO2–O2 (2:1) battery and Li/CO2 battery. Energy Environ. Sci. 7, 677 (2014). doi:10.1039/c3ee43318h

19. Z. Guo, X. Dong, S. Yuan, Y. Wang, Y. Xia, Humidity effect on electrochemical performance of Li–O2 batteries. J. Power Sources 264, 1–7 (2014). doi:10.1016/j.jpowsour.2014.04.079

20. Y. C. Lu, H. A. Gasteiger, Y. Shao-Horn, Catalytic activity trends of oxygen reduction reaction for nonaqueous Li-air batteries. J. Am. Chem. Soc. 133, 19048–19051 (2011). Medline doi:10.1021/ja208608s

21. B. D. McCloskey, R. Scheffler, A. Speidel, D. S. Bethune, R. M. Shelby, A. C. Luntz, On the efficacy of electrocatalysis in nonaqueous Li-O2 batteries. J. Am. Chem. Soc. 133, 18038–18041 (2011). Medline doi:10.1021/ja207229n

22. S. H. Oh, L. F. Nazar, Oxide catalysts for rechargeable high-capacity Li-O2 batteries. Adv. Energy Mater 2, 903–910 (2012). doi:10.1002/aenm.201200018

23. S. H. Oh, R. Black, E. Pomerantseva, J. H. Lee, L. F. Nazar, Synthesis of a metallic mesoporous pyrochlore as a catalyst for lithium–O2 batteries. Nat. Chem. 4, 1004–1010 (2012). Medline doi:10.1038/nchem.1499

24. J. Lu, Y. Lei, K. C. Lau, X. Luo, P. Du, J. Wen, R. S. Assary, U. Das, D. J. Miller, J. W. Elam, H. M. Albishri, D Abd EI-Hady, Y. K. Sun, L. A. Curtiss, K. Amine, Nat. Commun. 4, 2383 (2013).

25. H. G. Jung, Y. S. Jeong, J. B. Park, Y. K. Sun, B. Scrosati, Y. J. Lee, Ruthenium-based electrocatalysts supported on reduced graphene oxide for lithium-air batteries. ACS Nano 7, 3532–3539 (2013). Medline doi:10.1021/nn400477d

26. E. Yilmaz, C. Yogi, K. Yamanaka, T. Ohta, H. R. Byon, Promoting formation of noncrystalline Li2O2 in the Li-O2 battery with RuO2 nanoparticles. Nano Lett. 13, 4679–4684 (2013). Medline doi:10.1021/nl4020952

27. B. Sun, X. Huang, S. Chen, P. Munroe, G. Wang, Porous graphene nanoarchitectures: An efficient catalyst for low charge-overpotential, long life, and high capacity lithium-oxygen batteries. Nano Lett. 14, 3145–3152 (2014). Medline doi:10.1021/nl500397y

Page 42: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

3

28. J. Xiao, D. Mei, X. Li, W. Xu, D. Wang, G. L. Graff, W. D. Bennett, Z. Nie, L. V. Saraf, I. A. Aksay, J. Liu, J. G. Zhang, Hierarchically porous graphene as a lithium-air battery electrode. Nano Lett. 11, 5071–5078 (2011). Medline doi:10.1021/nl203332e

29. Z. L. Wang, D. Xu, J. J. Xu, L. L. Zhang, X. B. Zhang, Graphene oxide gel-derived, free-standing, hierarchically porous carbon for high-capacity and high-rate rechargeable Li-O2 batteries. Adv. Funct. Mater. 22, 3699–3705 (2012). doi:10.1002/adfm.201200403

30. H.-D. Lim, K.-Y. Park, H. Song, E. Y. Jang, H. Gwon, J. Kim, Y. H. Kim, M. D. Lima, R. O. Robles, X. Lepró, R. H. Baughman, K. Kang, Enhanced power and rechargeability of a Li-O2 battery based on a hierarchical-fibril CNT electrode. Adv. Mater. 25, 1348–1352 (2013). Medline doi:10.1002/adma.201204018

31. M. J. Lacey, J. T. Frith, J. R. Owen, A redox shuttle to facilitate oxygen reduction in the lithium air battery. Electrochem. Commun. 26, 74–76 (2013). doi:10.1016/j.elecom.2012.10.009

32. Y. Chen, S. A. Freunberger, Z. Peng, O. Fontaine, P. G. Bruce, Charging a Li-O₂ battery using a redox mediator. Nat. Chem. 5, 489–494 (2013). Medline doi:10.1038/nchem.1646

33. D. Sun, Y. Shen, W. Zhang, L. Yu, Z. Yi, W. Yin, D. Wang, Y. Huang, J. Wang, D. Wang, J. B. Goodenough, A solution-phase bifunctional catalyst for lithium-oxygen batteries. J. Am. Chem. Soc. 136, 8941–8946 (2014). Medline doi:10.1021/ja501877e

34. H. D. Lim, H. Song, J. Kim, H. Gwon, Y. Bae, K.-Y. Park, J. Hong, H. Kim, T. Kim, Y. H. Kim, X. Lepró, R. Ovalle-Robles, R. H. Baughman, K. Kang, Superior rechargeability and efficiency of lithium-oxygen batteries: Hierarchical air electrode architecture combined with a soluble catalyst. Angew. Chem. Int. Ed. 126, 4007–4012 (2014). doi:10.1002/ange.201400711

35. W. J. Kwak, D. Hirshberg, D. Sharon, H.-J. Shin, M. Afri, J.-B. Park, A. Garsuch, F. F. Chesneau, A. A. Frimer, D. Aurbach, Y.-K. Sun, Understanding the behavior of Li–oxygen cells containing LiI. J. Mater. Chem. A 3, 8855–8864 (2015). doi:10.1039/C5TA01399B

36. M. M. Ottakam Thotiyl, S. A. Freunberger, Z. Peng, Y. Chen, Z. Liu, P. G. Bruce, A stable cathode for the aprotic Li-O2 battery. Nat. Mater. 12, 1050–1056 (2013). Medline doi:10.1038/nmat3737

37. K. J. Hanson, C. W. Tobias, Electrochemistry of iodide in propylene carbonate. J. Electrochem. Soc. 134, 2204 (1987). doi:10.1149/1.2100852

38. B. D. Adams, R. Black, C. Radtke, Z. Williams, B. L. Mehdi, N. D. Browning, L. F. Nazar, The importance of nanometric passivating films on cathodes for Li-air batteries. ACS Nano 8, 12483–12493 (2014). Medline doi:10.1021/nn505337p

39. M. Leskes, N. E. Drewett, L. J. Hardwick, P. G. Bruce, G. R. Goward, C. P. Grey, Direct detection of discharge products in lithium-oxygen batteries by solid-state NMR spectroscopy. Angew. Chem. Int. Ed. 51, 8560–8563 (2012). doi:10.1002/anie.201202183

40. S. S. Zhang, D. Foster, J. Read, Discharge characteristic of a non-aqueous electrolyte Li/O2 battery. J. Power Sources 195, 1235–1240 (2010). doi:10.1016/j.jpowsour.2009.08.088

Page 43: Supplementary Materials for - Science€¦ · polarization due to the diffusion of electrode active species and ohmic loss caused by the insulating discharge product. The rGO electrodes,

4

41. L. Johnson, C. Li, Z. Liu, Y. Chen, S. A. Freunberger, P. C. Ashok, B. B. Praveen, K. Dholakia, J. M. Tarascon, P. G. Bruce, The role of LiO2 solubility in O2 reduction in aprotic solvents and its consequences for Li-O2 batteries. Nat. Chem. 6, 1091–1099 (2014). Medline doi:10.1038/nchem.2101

42. N. B. Aetukuri, B. D. McCloskey, J. M. García, L. E. Krupp, V. Viswanathan, A. C. Luntz, Solvating additives drive solution-mediated electrochemistry and enhance toroid growth in non-aqueous Li-O₂ batteries. Nat. Chem. 7, 50–56 (2015). Medline doi:10.1038/nchem.2132

43. W. S. Hummers Jr., R. E. Offeman, Preparation of graphitic oxide. J. Am. Chem. Soc. 80, 1339 (1958). doi:10.1021/ja01539a017

44. Y. Zhao, L. Wang, H. R. Byon, High-performance rechargeable lithium-iodine batteries using triiodide/iodide redox couples in an aqueous cathode. Nat. Commun. 4, 1896 (2013). Medline doi:10.1038/ncomms2907

45. L. H. Saw, Y. Ye, A. A. O. Tay, Electrochemical–thermal analysis of 18650 lithium iron phosphate cell. Energy Convers. Manage. 75, 162–174 (2013). doi:10.1016/j.enconman.2013.05.040

46. P. Stevens, G. Toussaint, G. Caillon, P. Viaud, P. Vinatier, C. Cantau, O. Fichet, C. Sarrazin, M. Mallouki, ECS Trans. 28, 1 (2010).

47. R. S. Assary, K. C. Lau, K. Amine, Y. K. Sun, L. A. Curtiss, Interactions of dimethoxy ethane with Li2O2 clusters and likely decomposition mechanisms for Li–O2 batteries. J. Phys. Chem. C 117, 8041–8049 (2013). doi:10.1021/jp400229n