Electrochemical Studies of Lithium-Oxygen Reactions for } ARCHGvES

advertisement
Electrochemical Studies of Lithium-Oxygen Reactions for
Lithium-Air Battery Applications
ARCHGvES
MASSACHUSES IN5TMJTE
d.
K
OF TECHNOLOGY
-
David G. Kwabi
-
JUN 2 5 2013
B.S.E. Mechanical Engineering
L B -RARI E S
Princeton University, 2010
SUBMITTED TO THE DEPARTMENT OF MECHANICAL ENGINEERING IN PARTIAL
FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF
MASTER OF SCIENCE IN MECHANICAL ENGINEERING
AT THE
MASSACHUSETTS INSTITUTE OF TECHNOLOGY
June 2013
0 2013 Massachusetts Institute of Technology
All rights reserved
...................
Signature of Author.............................
Department of Mechanical Engineering
} ay 23, 2013
Certified by....................
ang Shao-Horn
Gail E. Kendall Professor of Mechanical Engineering
Thes
Supervisor
Accepted by........................
David E. Hardt
Ralph E. and Eloise F. Cross Professor of Mechanical Engineering
Chairman, Department Graduate Committee
1
2
Electrochemical Studies of Lithium-Oxygen Reactions for Lithium-Air
Battery Applications
By
David G. Kwabi
Submitted to the Department of Mechanical Engineering
on May 10, 2013 in Partial Fulfillment of the
Requirements for the Degree of Master of Science in
Mechanical Engineering
ABSTRACT
Fundamentally understanding reaction mechanisms during Li-0 2 cell operation is critical for
implementing Li-air batteries with high reversibility and long cycle life. In this thesis, the
rotating ring disk electrode (RRDE) technique has been used to probe the influence of different
electrolyte solvents on the stability of the superoxide radical produced on planar glassy carbon
and Au electrodes. It was found that the fraction of oxygen reduction reaction current attributable
to superoxide generation exhibits a solvent-invariant potential dependence on carbon, with a
higher fraction of superoxide produced at lower discharge overpotentials. This trend is in support
of a proposed growth model for different Li 20 2 morphologies, where Li 2 0 2 growth is governed
primarily by disproportionation of superoxide at low overpotentials and direct electron transfer at
high overpotentials. On Au, superoxide stability exhibits a strong solvent dependence, which can
be explained in terms of the effect of the electrolyte solvent basicity on the stability of the Li+02~ ion pair.
This study highlights the potential use of RRDE as a tool to gain insights into Li-0 2 reaction and
growth mechanisms and the contribution of soluble intermediate species to parasitic reactions in
practical Li-air batteries.
Thesis Supervisor: Yang Shao-Horn
Title: Gail E. Kendall Professor of Mechanical Engineering
3
Acknowledgements
I would like to thank my adviser Professor Yang Shao-Horn for her constant support and
guidance throughout this work, and for continually encouraging me to think rigorously about
scientific questions. I am particularly grateful for how generous she was with her time, in
meeting with me in order to give useful feedback on my ideas and results.
I would also like to thank Yi-Chun Lu for taking the time to show me how to perform
electrochemical measurements, and for teaching me a lot about electrochemistry and how to
interpret results. I owe a lot to Betar Gallant for giving me pointers on how to motivate, frame
and present my work and for sitting through my practice presentations and providing comments
on my writing.
It has been and continues to be a pleasure to work with everyone else in the electrochemical
energy lab (EEL). I am amazed by how friendly and supportive people are here. I would
especially like to thank my officemates Alexis Grimaud and Wesley Hong for letting me bounce
ideas off them and giving me very good feedback, as a result of which I learned a lot about solidstate physics and chemistry. I am grateful to Joe Elias, who has taught me a lot about chemistry,
as well as Jonathon Harding and Koffi Yao for their excellent feedback and probing questions
during our weekly group meetings.
I would lastly like to thank my family: Papito, Bozi, Bella, Duiko and Obaa. I could not have
done this without their support. Many friends, near and far, have supported me in this endeavor,
and my thanks goes to them as well.
This work was supported by a Total Graduate Student Fellowship and the U.S. Department of
Energy's U.S.-China Clean Energy Research Center for Clean Vehicles (Grant DE-PI0000012).
5
6
Table of Contents
A cknow ledgem ents.........................................................................................................................
5
Table of Contents............................................................................................................................
7
List of Figures.................................................................................................................................
9
Introduction ...........................................................................................................................
11
1.
2
3
4
1.1.
M otivation......................................................................................................................
11
1.2
Operation and Challenges ..........................................................................................
13
1.3
Current U nderstanding of the Li-ORR Pathway ........................................................
18
1.4
M orphological characteristics of Li 202......................
1.5
Scope of W ork................................................................................................................
23
Experim ental M ethods...........................................................................................................
27
2.1
RRD E operation .............................................................................................................
27
2.2
Electrode Preparation .................................................................................................
29
2.3
RRD E Collection Efficiency Calibration....................................................................
30
2.4
Superoxide Collection Experim ents...........................................................................
32
.. . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Results ...................................................................................................................................
..
21
35
3.1
Collection Efficiency Calibration................................................................................
35
3.2
Rotation D ependence of Superoxide Collection ........................................................
36
3.3
Potential Dependence of Superoxide collection on Carbon.......................................
39
3.4
Potential D ependence of Superoxide Collection on Au..............................................
47
3.5
M echanistic Implications ............................................................................................
53
Conclusions and Future W ork .............................................................................................
57
References.....................................................................................................................................
61
7
8
List of Figures
Figure 1 Cathodic and anodic reaction potential ranges and corresponding gravimetric capacities
13
of selected lithium battery chem istries.I...................................................................................
Figure 2 Ragone plot showing gravimetric power and energy densities based on the cathode
weight for selected lithium-ion and Li-0 2 batteries. Li-0 2 cathodes considered are based on
freestanding hierarchically porous carbon (FHPC graphene) 7 , carbon nanofibers (CNF)8 , Vulcan
carbon (VC) 9 , Super P carbon 1 0 and pristine Nao 4 4 MnO 2 nanowires/Ketjen black (P-Z
15
M nO 2/K B ) . ..................................................................................................................................
ions
Li*
discharge,
During
cell.
a
Li-0
of
principles
working
illustrating
Figure 3 Schematic
2
from the anode travel through Li-containing electrolyte and combine with reduced oxygen at the
cathode to form Lix0 2 , where x is typically 2 (Li 2 0 2 ). The reverse process is expected to occur
17
o n ch arg e .......................................................................................................................................
Figure 4 CVs showing ORR and OER in 0.5M TBAC1O4 in DME and GC, polycrystalline Au,
Pt and Ru electrodes and in 0.1 M LiClO4 in DME (increased 5x for comparison with ORR and
OER in TBA) on glassy carbon at 20 mV/s.24
.....................................
19
Figure 5 Scanning Electron Microscope (SEM) images of (a) pristine Vulcan Carbon electrode
(b) Vulcan Carbon electrode discharged at 100 mAg' (to 2500 mAh gcarbon ) and (c) Vulcan
Carbon electrode discharged at 1000 mAg'1 (to 1400 mAhgcarbon )9..............................
........ 22
Figure 6: Schematic showing design of RRDE. Electroactive species of interest generated at the
disk electrode are radially convected (by rotation) to the ring electrode, where they can detected
27
(oxidized or reduced) by setting the ring at the appropriate potential......................................
Figure 7 Schematic of section view of RRDE, illustrating its operation. Reduction of 0 at the
disk is followed by oxidation of R at the ring. Arrows indicate direction of electron flow. ........ 29
Figure 8 Rotating disk electrode setup showing working, reference and counter electrodes.
F igure C redit: Yi-C hun L u............................................................................................................30
Figure 9 Flowchart describing experimental and analytical protocols for superoxide collection
33
a ssessm ent.....................................................................................................................................
Figure 10 Steady-state cyclic voltammograms of ferrocene oxidation/ferrocenium reduction on
35
glassy carbon and Au at 0 rpm in 50mM Fc in 0.1M LiClO 4 in DME..................
Figure 11 Disk and ring currents of the Fc/Fc* redox reaction on (a) glassy carbon and (b) Au at
400, 600 and 900 rpm in 0.5mM ferrocene in 0.1M LiClO 4 in DME. The disk was held at 4.0 V
while the ring was held at 3.0 V vs Li/Lie to ensure diffusion-limited oxidation of Fc and
reduction of Fc. The collection efficiency, N, on glassy carbon was 25.80 ± 0.32%, while that
for Au was 25.97 ± 0.47%. Error bars were calculated from the standard deviation of values
36
obtained at 400, 600 and 900 rpm .............................................................................................
1600
900,
Figure 12 (a) CVs of ORR/OER in 0.1M LiClO 4 in DME on glassy carbon at 0, 400,
and 2700rpm showing disk (left axis) and ring (right axis) current densities. Ring is held at 3.5 V
(b) Estimated fractions of LiO 2 (oxidized on ring) and Li 2 0 2 (oxidized on disk) formed from
37
total O RR ch arge...........................................................................................................................
2,2
2.6,
2.4,
at
in
DME
M
LiClO
0.1
Figure 13 (a) Disk current transients in Ar and 0 2-saturated
4
and 2.0 V (b) Background-subtracted disk charge (c) Background-subtracted ring charge and (d)
Ring charge/Disk charge versus applied potential in 0.1M LiClO 4 in DME (blue), 0.1M LiClO 4
in DMSO (red) and 0. 1M LiClO 4 in DMF (green) on glassy carbon. RRDE rotation was at 900
41
rp m ................................................................................................................................................
9
Figure 14 Cyclic voltammograms for ORR and OER in (a) 0.1 M LiPF 6 in DME and (b) 0.1 M
LiPF 6 in DMSO at various potential windows on glassy carbon at 100 mV/s 5 ..................... 45
Figure 15 (a) Disk current transients in Ar and 0 2 -saturated 0. 1M LiClO 4 in DME at 2.6, 2.4, 2,2
and 2.0 V (b) Background-subtracted disk charge (c) Background-subtracted ring charge and (d)
Ring charge/Disk charge versus applied potential in 0.1M LiClO 4 in DME (blue) and 0.1M
LiClO 4 in DM SO (red) on Au at 900rpm . ................................................................................
48
Figure 16 Ring charge/Disk charge versus applied potential in 0.1 M LiClO 4 in DMSO on carbon
(filled circles) and Au (open circles) at 900rpm. Error bars are calculated from 3 independent
m easurem ents................................................................................................................................
50
Figure 17 Linear sweep voltammetry measurements at 50 mV/s following potentiostatic ORR at
2.7, 2.6, 2.4 and 2.0 V on (a) glassy carbon in 0.1M LiClO 4 in DME (b) Au in 0.1M LiClO 4 in
DME (c) glassy carbon in 0.1M LiClO 4 in DMSO (d) Au in 0.1M LiClO 4 in DMSO to oxidize
electrodeposited species................................................................................................................
51
Figure 18 Separate measurements of ring charge/disk charge versus applied potential in 0.1M
LiClO 4 in DMSO, DMF and DME on (a) glassy carbon and (b) Au .......................................
52
10
1. Introduction
1.1.
Motivation
Over the past three decades, the scientific consensus that increased fossil fuel consumption is
leading to irreversible climate change has solidified.' This realization, coupled with the depletion
of fossil resources has generated interest in wind, solar, tidal and geothermal power as future
energy sources. These energy sources are however irregular in nature, and therefore must be used
within a framework of comprehensive energy storage, primarily via batteries and chemical fuels
such as hydrogen and biofuels.
One major renewable energy need is in the area of transportation. 28% of all fossil fuel usage is
directly related to transport2 , and with the growth of the global economy, increased demand for
automobiles threatens to escalate global CO 2 emissions even further. Finding adequate energy
storage alternatives to fossil fuels for electromotive applications should therefore be a key
component of any future clean energy infrastructure.
Batteries are attractive as energy storage devices due to their portability and suitability to a wide
variety of portable electronic applications. Lithium-ion batteries in particular have been
successfully used in a wide range of devices such as phones, laptops and remote sensors. Despite
their relatively effective commercialization, most lithium-ion battery configurations have low
1 http://www.ipcc.ch/publicationsanddata/ar4/syr/en/mains2-4.html
Institute for Energy Research Report. Hard Facts: An Energy Primer
http://instituteforenergyresearch.org/hardfacts.pdf)
2
(online at
11
energy storage capabilities and would be too costly for more energy-intensive applications such
as electric grid load leveling and fully electric vehicles.
The gravimetric energy density (in Wh per kilogram of active material weight) of any battery
chemistry can be calculated based on the potential difference (in volts) between anodic and
cathodic half-reactions, and the amount of charge stored per unit weight of material (or capacity,
in mAh/g).
Figure 1 plots the reaction potentials and associated gravimetric capacities of
selected battery half-reactions.' The gravimetric energy density of each configuration equals the
product of the gravimetric capacity and overall cell reaction potential, i.e. the potential difference
between cathodic and anodic redox levels. State-of-the-art lithium-ion cathodes such as LiCoO
2
and LiNio. 5Mno.50 2 have theoretical energy densities ranging between 500 - 600 Wh/kgcathode, but
practical energy densities (including the weight of other cell package components such as
electrolyte, separator and binder) of 30 - 40% of that value, and would be impractical for electric
vehicle applications, which require batteries with energy density of at least 500 Wh/kg at the
package level. 2 ,3
Lithium-oxygen (Li-0 2) batteries, on the other hand, are projected to store 1500 - 2000
Wh/kgcathode,4 and are thus attracting increasing attention as future energy storage units for
automotive applications. This thesis will focus exclusively on using electrochemical techniques
to fundamentally understand and thereby improve energy storage capabilities in prototypical
non-aqueous Li-0
2
batteries.
12
r
Podacton cmpotntds
9O)A44AM-CCo ..
1.2
2per
tn
a
omposi
ait
Li02bateed
1
C
l
ce
a
UvFiPm
I
ng0ga
c
3d-Mle oxn
iSnO)bai
in"-bBued
Capaciy (AhkW')
Figure 1 Cathodic and anodic reaction Ipotential ranges and corresponding gravimetric capacities
of selected lithium battery chemistries.1
1.2
Li-0 2
Operation and Challenges
batteries differ from traditional Li-ion batteries in that rather intercalating into a host
lattice, Li ions react directly with oxygen from air dissolved in the electrolyte, resulting in the
formation of lithium-oxygen compounds. In rechargeable non-aqueous Li-0 2 batteries, Li* ions
are formed from the oxidation of 'netallic Li at the anode and travel through the electrolyte to the
cathode during discharge. At the cathode, is oxygen is simultaneously reduced in the presence of
these Li+ ions to form a solid, insoluble Li 2O2 phase (Figure 3), while electrons in the external
circuit perform electrical work. This process is reversed during charge with the decomposition of
Li 20 2 , evolution of molecular oxygen and plating of Li at the anode.
This open framework of operation dispenses with the need to use heavy transition metal-based
cathodes, which are typically employed in intercalation structures in Li-ion batteries. The loss of
these transition metals and their replacement with lightweight materials (such as carbon) as
13
positive electrodes is expected to contribute to the high energy density of Li-0 2 batteries.
Additionally, since the capacity of the battery is theoretically limited by the filling of void
volume in the cathode, the combination of a high theoretical capacity of Li 20 2 (1168 mAh/gu20 2 )
and highly porous cathodes is expected to boost these gains even further.
Much research over the last two decades has gone into the development of Li-0
2
positive
electrodes that can realize the projected gravimetric energy density gains over lithium-ion
batteries. Gravimetric power and energy densities based on the cathode weight for selected
lithium-ion and Li-0 2 cathodes are shown in the Ragone plot in Figure 2.4 At a gravimetric
power of ~100 W/kg, lithium-ion cathodes such as LiCoO 2 5 and LiNiO. 5Mn. 50 26 can deliver
energy densities between 500 - 800 Wh/kg, while various Li-0
2
cathodes based on graphene,
carbon nanofibers, and Vulcan carbon have demonstrated 1800 - 2800 Wh/kg. These energy
densities suggest that Li-0 2 batteries provide a route to achieving 3 - 5 times the energy density
of current lithium-ion battery configurations.
14
104
I UNI
UI
5MnO 2
02
FHPC
Graphene
CNF
103
UFePO
10 ~\vc
102
LI0o0, Super
-ZMnO IKB
10
101
102101
,)
Gravimetric Energy (Whkg
Figure 2 Ragone plot showing gravimetric power and energy densities based on the cathode
weight for selected lithium-ion and Li-0 2 batteries. Li-0 2 cathodes considered are based on
freestanding hierarchically porous carbon (FHPC graphene)7, carbon nanofibers (CNF)8 , Vulcan
carbon (VC) 9 , Super P carbon" and pristine Nao. 44MnO2 nanowires/Ketjen black (P-Z
.
MnO 2/KB)'
Although laboratory-scale lithium-oxygen cells have demonstrated gravimetric energy densities
that are -several-fold higher than those of current Li-ion batteries7'
level, Li-0
2
' '
at the positive electrode
battery technology still faces several device-level challenges for practical
application, such as high charging overpotentials, low rate capability and poor cycle life.
One of the major fundamental impediments to such practical application is the inability to find
an electrolyte that is chemically stable to oxygen reduction reagtion (ORR) intermediates and
products.
In early studies, organic carbonate-based electrolyte solvents commonly used in
lithium-ion batteries were used in electrolytes for prototypical Li-0 2 cells. They have however
been shown to react with the superoxide (02~) ORR intermediate 14-16, resulting in the formation
of side products such as Li 2 CO 3 , HCO 2Li and CH 3 CO 2Li duing discharge. These species were
15
found to be difficult to decompose during charge and to accumulate during electrochemical
cycling, leading to observed capacity losses and high charging overpotentials. Despite the
widespread transition to relatively more stable ether-based solvents, where Li2 0 2 is the principal
discharge product, electrolyte decomposition and capacity fade continue to persist10 , presumably
as a result of ether autooxidation in the presence of 02.17 Silane and amide-based solvents have
also been evaluated for use as electrolyte solvents in Li-air batteries. While an oligoetherfunctionalized silane solvent exhibited improved stability compared to organic carbonates, some
electrolyte degradation was observed by Fourier-Transform Infrared Spectroscopy (FTIR).18
Recent reports have suggested a high degree of direct chemical reactivity between the superoxide
radical or the final discharge product Li 2 0 2 and carbon to form Li 2 CO3 19 ,2 0 , which then
contributes further to the high charging overpotentials and poor cyclability seen in carbon-based
Li-0
2
cells. Additionally, the superoxide radical has also been shown to react with other cell
components such as salt and binder. Chemically generated superoxide has been observed to
readily dehydrofluorinate the common binder material polyvinylidene difluoride (PVDF)2 1
resulting in the formation of fluorine-related side products. Similar species have identified by
XPS in Li-0
2
cathodes discharged in electrolytes with non-fluorine-containing salts.1 2 XPS has
been further used to identify halide species related to the degradation of anions in LiBF 4 , LiTFSI,
LiPF6 and LiCIO 4 salts used in Li-0
2
cathodes2 2 suggesting that Li2 0 2 is slightly unstable in the
presence of most lithium salts currently used in electrolyte configurations for nonaqueous
lithium-oxygen batteries.
16
Negative Electrode
Positive Electrode
Figure 3 Schematic illustrating working principles of a Li-0 2 cell. During discharge, Li+ ions
from the anode travel through Li-containing electrolyte and combine with reduced oxygen at the
cathode to form Lix0 2 , where x is typically 2 (Li 2 0 2). The reverse process is expectedto occur
on charge.
Due to the scale and complexity of possible chemical interactions between various cell
components and electroactive
species/intermediates
involved in Li-0
2
electrochemistry,
understanding the identity, stability and degree of solubility of Li-0 2 reaction intermediates
under various environmental conditions is a critical task. In particular, such an effort will be key
to identifying reaction mechanisms responsible for the formation and growth of Li 2 O2 , but also
for assessing the effect of reaction intermediate stability on discharge product chemistry and the
aforementioned device-level performance challenges such as low roundtrip efficiency and poor
cycle life.
17
1.3
Current Understanding of the Li-ORR Pathway
The ORR mechanism on some noble metal electrode and carbon surfaces in nonaqueous
electrolyte solvents with tetra-alkylammonium-containing salts has been fairly well established
from electrochemical techniques.1 4 2 3 Oxygen reduction proceeds by highly reversible letransfer to molecular oxygen at ~ 2.0 V vs Li/Li* (all voltages hereafter referred to as 'V'
referenced to Li/Lie redox), forming stable, soluble superoxide species which are then solvated
by the aforementioned cations (Figure 4). An example of such a reaction with the
tetrabutylammonium cation is as follows:
TBA+ +02+ e- " TBAO 2
(1)
A similarly detailed understanding of the ORR in the presence of Li+ ions has proven more
elusive. This is due to the relative chemical instability of the O2/LiO 2 species, which is presumed
to be the first intermediate formed upon ORR (analogous to reaction 1, with Li* replacing
TBA*). This species can react with cell components as discussed above, or undergo a chemical
disproportionation reaction with other LiO 2 radicals to produce a solid-phase Li 2 0 2 , which
precipitates on and passivates the electrode, inhibiting any further reaction.
18
2
GC
EE
SIM1 Au
0a+2'+ 2LJ*-O LUP
GC
Pt
1.0
1.5
2.0
2.5
3.0
3.5
4.0
4.5
E (V vs. Li)
Figure 4 CVs showing ORR and OER in 0.5M TBAC1O4 in DME and GC, polycrystalline Au,
Pt and Ru electrodes and in 0.1 M LiC1O4 in DME (increased 5x for comparison with ORR and
OER in TBA) on glassy carbon at 20 mV/s.24
The first direct observation of superoxide-related reaction intermediates during Li-ORR was
reported by Peng et al 5 who studied the Li-ORR with in situ Raman spectroscopy. Upon
holding an Au electrode at an applied bias of 2.2 V vs Li/Li* in an acetonitrile-based electrolyte,
a Raman peak attributable to the 0-0 stretching frequency in Li0 2 was observed at 1137 cm'
After ~ 4 min of sustained polarization, a new peak corresponding to the 0-0 stretch in Li 2 0 2
emerged at 808 cm-' at the expense of the original superoxide peak. The sequential formation of
Li 2 0 2 after Li0 2 suggests that one possible mechanism for the formation of Li 2 0 2 is analogous to
proton-induced disproportionation of HO 223 , where the formation of Li 20
2
proceeds via Li*-
induced chemical disproportionation of LiO 2 :
2LiO2.-> Li 2 02+ 02
(2)
19
Cyclic voltammetry measurements obtained using the rotating ring disk electrode (RRDE)
technique also suggest that LiO 2 has limited solubility in some electrolyte solvents such as
dimethyl sulfoxide (DMSO)
26,27.
This is inferred from the positive ring currents observed during
ORR on the disk electrode while the ring electrode is held at a sufficiently high potential (~ 3 3.5 V vs Li/Lie) to oxidize superoxide-related species convected from the disk surface. These
two observations together result in one possible Li 2 0 2 formation reaction mechanism where the
generation of LiO 2 corresponds to a process analogous to that of step 1 followed by step 2.
The persistence of stable and highly nucleophilic 02~-related species during ORR has been
deemed responsible for the degradation of electrolyte solvents 0 , salt 12 and binder2,21, which has
been shown to result in poor cycle life in Li-0 2 cells, as previously discussed.
While steps 1 and 2 comprise a plausible reaction mechanism for Li 2 0 2 formation, there is
evidence from literature for a different mechanism, where step 1 is followed by electron transfer
to the superoxide species:
Li+ + LiO 2 + e -> Li 2 0
2
(3)
Recent differential electrochemical mass spectrometry (DEMS) measurements during linear
sweep voltammetry and galvanostatic discharge in ether and DMSO-based electrolytes have
given evidence that ~2e-/mole of 02 is consumed during Li-ORR on carbon in DME2 8 . While
this result is consistent with the formation of LiO 2 (step 1) followed by very fast chemical
disproportionation (step 2), the authors report no evidence for a le- reduction of 02 to form 02~
/LiO 2 -type species at any ORR potentials at or below 2.75 V, as would be expected from
spectroscopic measurements 25 . A consecutive electron transfer step such as that depicted in step
20
3 was therefore deemed by the authors as more likely to be the dominant reaction mechanism in
the formation of Li 2 O2 .
Thus, while the overall reaction attributed to the formation of Li 2 0 2 is 2Li + 02
+
Li2 0 2 , two
different pathways for the formation of Li 20 2 are possible: upon the generation of LiO 2 in step 2,
Li 20 2 formation may proceed via chemical disproportionation in step 3 or consecutive electron
transfer to LiO 2 in step 4. Well-defined investigative schemes where the kinetics of the Li-0
2
reaction and the stability of LiO 2 can be simultaneously monitored as a function of the electrolyte
and reaction surface used as well as applied potential are therefore critical for identifying a
generalizable reaction mechanism.
1.4
Morphological characteristics of Li 2O 2
In addition to the role of intermediates on Li 2 0 2 formation and parasitic reactivity, questions
remain regarding the effect of LiO 2 lifetime on the growth mechanisms of Li 20 2, and in particular
on the various discharge product morphologies that have been observed in Li-0
2
cathodes. In
particular, large toroids and disc-shaped features (see Figure 5b) have been observed in various
carbon-based cathodes.7,8,i, 29 While the growth of mechanism of these toroidal features are not
yet completely understood, they represent a possible route to efficiently filling the void volume
of the cathode and hence result in high energy density Li-0 2 batteries, in contrast to thin-film
deposits of Li 2 0 2, which poorly utilize the available pore space in the cathode.
21
(
(c)
n
Figure 5 Scanning Electron Microscope (SEM) images of (a) pristine Vulcan Carbon electrode
(b) Vulcan Carbon electrode discharged at 100 mAg1 (to 2500 mAh gearbon~1) and (c) Vulcan
Carbon electrode discharged at 1000 mAg' (to 1400 mAhgebaron~)
9
Several studies have been devoted to understanding nucleation, growth and decomposition
mechanisms of these toroidal morphologies, which range from several hundred nanometers to ~
Ipm in size. 29-31 These studies have been particularly motivated by the apparent discrepancy
between these feature sizes and the fact that Li 2 0 2 is a bulk insulator with a band gap of between
4 - 5 eV32-3 4, and has been shown to be limited to a thickness of 5 - 1Onm when
electrochemically grown on planar electrodes. 3 5 ,36 Recent studies have suggested that these two
morphologies of Li 20 2 exhibit a discharge rate/overpotential dependence, with toroids forming at
low applied ORR overpotentials (> 2.74 V) and thin deposits at large applied potentials (< 2.6V)
and current densities.
30,37
It has additionally been shown from galvanostatic tests3 0' 37 and the
potentiostatic intermittent titration technique 37 that the toroidal morphologies have higher
charging potential potentials than the thin-film deposits of Li 2 0 2 . This difference in Li 2O2
decomposition overpotential raises the prospect of fundamental surface chemical/electronic
differences between these different structures.
To probe this possibility, surface sensitive X-ray absorption near edge structure (XANES)
measurements of the different morphologies were made, and suggest that disc-shaped Li 2 0 2 have
22
superoxide-like
surface chemistry, while particles are largely stoichiometric. 3 7 Such a
discrepancy in surface chemistry is in agreement with the hypothesis that disc-like Li 20
2
is
formed by the agglomeration and subsequent disproportionation of superoxide (steps 1 and 2),
while particulate Li 20 2 grows largely via direct, surface mediated electron transfer to LiO 2 (steps
1 and 3).
Systematically investigating the dependence of LiO 2 lifetime on applied potential, as well as the
nature of the electrolyte solvent and reaction surface is therefore critical to understanding its
contribution to the formation of Li 20 2 and/or parasitic products.
1.5
Scope of Work
This thesis reports an investigation of superoxide formation kinetics and stability by means of the
rotating ring disk electrode (RRDE) technique, using carbon and Au as reaction surfaces and 1,2dimethoxyethane (DME), dimethylformamide (DMF) and DMSO as electrolyte solvents. RRDE
is a technique that can be used for studying electrode reactions coupled to homogeneous
reactions, where some knowledge about disk electrode reaction products can be inferred by
measuring current at the annular ring. It is thus a suitable technique for monitoring the
relationship between the kinetics of LiO 2 formation on various potential cathode reaction
surfaces, and its subsequent stability in a variety of electrolyte solvents.
Carbon was chosen as a reaction surface because it is the most widely studied cathode surface in
prototypical Li-0
2
batteries. 3 ,4 ,25 ,38 Carbon is particularly attractive for its low cost, high
conductivity, low density and availability in a large number of highly porous architectures (see
Figure 5a). Such properties are highly desirable for Li-0 2 cathodes intended for energy-intensive
but affordable applications, where an electronically conducting scaffold would be required for
23
efficient 02 reduction, and a high void space necessary for high energy density. Additionally, Au
was chosen as a reaction surface because, with the exception of low cost and density, it possesses
similar qualities, but represents a potential improvement over carbon for its potentially higher
chemical stability against the superoxide radical compared to carbon. Thus, nanoporous Au is
attracting increasing attention as a Li-0 2 cathode13,39 and was recently shown for the first time to
promote long-term reversible cycling of a Li-0 2 battery. 13
DME has received substantial interest as an electrolyte solvent because: (1) in contrast to organic
4
carbonate-based electrolytes, DME is moderately stable against the superoxide radical' and (2)
it has been shown to have a very high oxygen solubility4 0 and excellent oxygen transport
9
properties at high discharge rates .
DMF was chosen because, like DME, it has been shown to be reasonably stable toward the
superoxide radical 2 3 ,4 1 ,4 2 in the presence of tetraalkylammonium cations, and was recently
evaluated as an electrolyte solvent for Li-air batteries. 39
DMSO was selected because it was recently shown to promote 95% capacity retention over 100
cycles when used as electrolyte solvent with a nanoporous Au cathode.' 3 Interestingly, such
13
39
capacity retention was not realized with analogous Au cathodes in DMF and DME -based
electrolytes, where more extensive electrolyte decomposition occurred during ORR.
This study aims to understand the relationship between Li-0 2 reaction intermediate stability and
the nature of the electrode/electrolyte by systematically studying intermediate reactivity for each
possible electrode-electrolyte couple among the parameters chosen. By so doing, it becomes
possible to decouple the unique influence of each reaction surface and electrolyte solvent on the
24
stability of these reaction intermediates, and their subsequent effect on the electrochemical
reversibility of the Li-0 2 reaction.
25
26
2 Experimental Methods
2.1
RRDE operation
In a conventional rotating disk electrode (RDE) setup, the kinetics of an electron transfer
reaction (for instance, the conversion of the electroactive species, 0, to the product R via a
heterogeneous electron-transfer reaction 0 + e -> R) can be readily inferred for a given surface.
Conceptually, this is achieved by controlling the potential bias applied to the electrode at which a
reaction of interest occurs (known as the working electrode), and measuring the corresponding
current response at different rotation speeds. RDE is a particularly attractive technique because
the steady state velocity profile in the vicinity of the electrode can be rigorously, calculated from
hydrodynamic equations. 4 3 This allows the prediction of seady
state diffdsion-limited
currents at
known rotation speeds and the inference of kinetic parameters by extrapolation to effectively
infinite rotation rates (for instance, in Koutecky-Levich analysis).4 3
rAu
Carbon dis
Figure 6: Schematic showing design of RRDE. Electroactive species of interest generated at the
disk electrode are iadially convected (by rotation) to the ring electrode, where they can detected
(oxidized or reduced) by setting the ring at the appropriate potential.
27
Since the RDE features a single electrode surface, however, information about the products
swept away by rotation cannot be easily discerned. The insertion of an independent annular ring
electrode in the rotating ring disk electrode configuration (Figure 6) overcomes this limitation, as
soluble species generated on the disk are pushed toward the ring by the radial flow of the fluid
induced by rotation.
By simultaneously measuring the current on the ring during the electrode reaction on the disk, it
is possible to measure properties of species generated at the disk surface. For example, by
measuring the ring current while holding it at a fixed potential where R is oxidized back to 0,
and simultaneously keeping the disk at at a potential so as to reduce 0 to R (Figure 7), the
number of R species successfully detected at the ring compared to the that generated at the disk
can be estimated for a series of different rotation speeds. Examining this ratio (ring current/disk
current) as a function of rotation speed can provide an estimate of the reaction rate between R
with various electrolyte solvents (since if a such a reaction exists, the ring current/disk current
ratio will be lower than expected for a nonreactive species). A similar approach has been used to
estimate the first-order rate constant of the reaction between superoxide and propylene
carbonate.
44
28
Au
Carbon disk
ring
O0'
R //r
Figure 7 Schematic of section view of RRDE, illustrating its operation. Reduction of 0 at the
disk is followed by oxidation of R at the ring. Arrows indicate direction of electron flow.
2.2
Electrode Preparation
Electrochemical measurements were performed in glass three-electrode cells (Chemglass, USA).
Working electrodes consisted of polycrystalline gold (Au) and glassy carbon (GC) (0.196 cm 2
disks; Pine, USA) surfaces, surrounded by a gold ring with 6.5 mm internal diameter and 7.5 mm
external diameter. All working electrodes were polished to a 0.05 gm mirror-finish, ultrasonicated in deionized water (18.2 M2 -cm, Millipore) for 5 min and vacuum-dried at 75 0 C for 8
hours before each experiment. All electrodes were kept in the vacuum oven and directly
transferred to a water-free glovebox (H20 < 0.1 ppm, Mbraun, USA) without exposing to the
ambient.
The three-electrode cell used for RRDE measurements donsisted of a lithium-foil counter
electrode and a bulk disk as the working electrode. The reference electrode consisted either of a
29
lithium foil or a silver wire immersed in 0.1M TBAClO 4 (Sigma Aldrich) and 0.01M AgNO 3
(BASi) in acetonitrile, which was calibrated against Li immersed in the chosen electrolyte.
RRDE experiments were performed in a water-free glovebox. Electrolyte solvents used were 0.1
M LiClO 4 in DME (BASF, USA), 0.1 M LiClO 4 in DMSO (Novolyte, USA) and 0.1M LiClO
4
in DMF (Sigma Aldrich)
Electrolyt
Conter Electrode
Figure 8 Rotating disk electrode setup showing working, reference and counter electrodes.
Figure Credit: Yi-Chun Lu
2.3
RRDE Collection Efficiency Calibration
In this thesis, the RRDE is primarily used to oxidize, or collect, soluble intermediate species
generated during oxygen reduction/evolution at the disk by holding the ring at the appropriate
potential. The expected diffusion-limited ring current for an analogous disk current during
rotation can be estimated by solving convective-diffusion equations with a fluid concentration
profile set such that, under steady-state conditions, the concentration of reactant species at the
disk and ring surfaces respectively is zero. The expected diffusion-limited ring current for a
30
given disk current can then be easily calculated for any RRDE geometry by solving these
equations. The ratio of the ring current, i,. to the disk current, id, is known as the collection
efficiency, N, of the electrode:
(4)
ir
For an RRDE of disk radius ri, and inner and outer ring radii of r2 and r3 respectively, N is
independent of the concentration and transport of properties of the reactants, and is described by:
N =1
- F
+ # 222{[(
[1l - F(a)] - (1 + a + #)1-F
N -(aj)+
3.
wherea=
2-
=
=
F(l) =
(
3
(1+ ]a + #
(5)
5
3)2/3
,3
_1,
a)
_
and the function F is defined by
r2
ln
/(1
3
+
/3)3
)1
+ -
3
tan
1
261/3 _ 1
31/2
+
1
Following a similar approach to that of Paulus et al45, the intrinsic RRDE collection efficiency
was calibrated with the ferrocene (Fc)/ferrocenium (Fc*) redox couple, where diffusion-limited
Fc oxidation disk currents were compared to analogous Fc* reduction ring currents, with the ring
set at 3.0 V, at 400, 600 and 900 rpm. Au and glassy carbon electrodes were prepared as
described above and the electrolyte was 0.5mM Fc in 0. 1M LiClO 4 in DME.
31
2.4
Superoxide Collection Experiments
To investigate the effect of rotation on superoxide collection, cyclic voltammograms (CV) were
obtained between 2 - 4.5 V and 400 - 2700 rpm in 0 2-saturated electrolyte, with the ring held at
3.5 V. Working electrodes were first prepared by the procedure described above in the electrode
preparation section. The working electrode was immersed into an Ar-purged electrolyte for 20
minutes prior to each cyclic voltammetry or chronoamperometry experiment. After steady-state
CVs were obtained in Ar (2.0 - 4.5 V, 50 mVs 1 ), the cell was purged with 02 for 20 min and
similar CVs were obtained in 0 2 -saturated electrolyte at 50 mVs' at 400, 900, 1600 and 2700
rpm while the ring was held at 3.5 V.
In order to systematically investigate the effect of applied potential on the formation of
superoxide (i.e. as a fraction of total ORR charge), collection experiments were performed as
follows: the disk potential was stepped from open circuit voltage (OCV) of 3.0 - 3.2 V to
selected potentials between 3.0 and 2.0 V (2.9, 2.8, 2.7, 2.6, 2.5, 2.4, 2.2 and 2.0 V) at 900 rpm
for 3 min at each potential, while the ring was held at 3.5 V. The procedure was repeated in 02saturated electrolyte, with the ring again held at 3.5 V to oxidize superoxide produced during
oxygen reduction.
For the Au disk, potential cycling was performed between 3.1 - 3.15 V during Ar and 02 purging
to inhibit the formation of hydroxide species' 4 . Additionally, for chronoamperometric
experiments in 0 2-saturated electrolyte, linear sweep voltammetry was performed between 3.0
and either 4.0 V (when using an Au disk) or 4.5 V (when using glassy carbon disk) after each 3
min hold at ORR potentials below 2.8 VU to oxidize any solid Li-0 2-related species deposited
32
during electrolysis. Capacitive correction of disk and ring currents was done by subtracting the
current measured under Ar from that found in pure 02 under identical potentiostatic conditions.
Figure 9 Flowchart describing experimental and analytical protocols for superoxide collection
assessment
33
34
3 Results
3.1
Collection Efficiency Calibration
Figure 10 shows CVs obtained at 50mV/s between 3.0 and 4.0 V and demonstrate a quasireversible Fc/Fc+ couple with a peak separation of -300mV on both Au and glassy carbon. To
ensure that Fc' reduction proceeded under diffusion-limited conditions for the collection
experiments, a potential of 3.0 V were chosen for the ring. The collection efficiency, N, on glassy
carbon was 25.80 ± 0.32% (Figure 1la) while that for Au was 25.97 ± 0.47% (Figure 1lb).
These values were in good agreement with the application of Equation 5 to the RRDE geometry
used in subsequent experiments (disk radius = 5mm, inner ring radius = 6.5mm, outer ring radius
=
7.5mm), which resulted in a geometric collection efficiency of 25.5%.
Fc -+ Fc++'
30
20
Au
10
Glassy Carbon
0
1
0
-1
-20
Fc++e +Fc
-301
2.8
.
a
3.0
.
3.2 3.4 3.6 3.8 4.0
Potendal (V vs LIILi)
4.2
Figure 10 Steady-state cyclic voltammograms of ferrocene oxidation/ferrocenium reduction on
glassy carbon and Au at 0 rpm in 50mM Fc in 0. lM LiClO 4 in DME.
35
(a)
10()
iGlsy Cabn
10
(b)<
10. Au(b
0
0
-10.
-30
-30
,OWmm
00rpm
40
.40
150-
120
600rpm
120
60M
0w
400rpm
go.
Fc -+ Fc++e
90
+
Fc-+FcW
..-
W
o0-
30 -30
0
0
2.8
3.0
3.2
3.4
3.6
3.8
4.0
Potentia
(Vva VU)
4.2
4.4
4.6
2.8
3.0
3.2
3.4
3.6
3.8
4.0
4.2
4.4
4.6
Potendal (Vvs U/U)
Figure 11 Disk and ring currents of the Fc/Fc redox reaction on (a) glassy carbon and (b) Au at
400, 600 and 900 rpm in 0.5mM ferrocene in 0.1M LiClO 4 in DME. The disk was held at 4.0 V
while the ring was held at 3.0 V vs Li/Lie to ensure diffusion-limited oxidation of Fc and
reduction of Fc+. The collection efficiency, N, on glassy carbon was 25.80 ± 0.32%, while that
for Au was 25.97 ± 0.47%. Error bars were calculated from the standard deviation of values
obtained at 400, 600 and 900 rpm.
3.2
Rotation Dependence of Superoxide Collection
Figure 12a shows CVs representing the ORR and oxygen evolution reaction (OER) in 02saturated 0. 1M LiClO 4 in DME between 2.0 - 4.5 V at rotation speeds between 0 and 2700 rpm.
At all rotation speeds, there is a monotonic increase in ORR current during the cathodic scan,
with an onset at -2.75 V, and a limiting ORR current around 0.24 pA/cm 2 . The presence of a
limiting peak, rather than a plateau, can be attributed either to charge transport limitations
through an insulating Li 2 0 2 layer deposited during the cathodic scan 14'46 or some ftmdamentally
36
kinetic limitation.4 7 This interpretation is supported by the fact that the ORR peak does not
increase with rotation4 7 , as would be expected for Li* or 02 mass transport-limited current.
No ring current is detected under static conditions, as no convective flow of superoxide
intermediates to the ring would occur. At 400, 900, 1600 and 2700 rpm, there is positive ring
current during ORR on the disk and a concomitant decrease in the peak OER current during the
anodic scan. This phenomenon is consistent with the convection and oxidation of a superoxidetype species during ORR, and hence the formation (and subsequent oxidation) of less Li2 0 2 than
would be expected under static conditions. The negligible ring current during anodic scans (3.0 4.0 V) upon rotation suggests that Li 20 2 oxidation proceeds without the involvement of soluble
superoxide radicals as during ORR, and is in agreement with in situ spectroscopic measurements
where LiO 2 was not observed during Li 2 0 2 oxidation 25 and with DEMS studies showing 2e/mole of 02 evolved during Li20 2 charging.13 28
0.2
(a)
50.n.t
.
d
io, 7k
E 04
100
e
(b)
Total
rpm
.r.1 4t rpm
uo
o~o
1600 rpm
2700 rpm
2.0
2.6
30
3U
40
Potential (ve LULl)
4
0
so
100
IN
2M
2W
30o
Rottm (rpm)
Figure 12 (a) CVs of ORR/OER in 0.1 M LiClO4 in DME on glassy carbon at 0, 400, 900, 1600
and 2700rpm showing disk (left axis) and ring (right axis) current densities. Ring is held at 3.5 V
(b) Estimated fractions of LiO2 (oxidized on ring) and Li2 0 2 (oxidized on disk) formed from
total ORR charge
37
It is interesting to note that ring and disk current transients, and hence collection efficiencies for
superoxide, are similar for all positive rotation speeds. This suggests firstly that similar amounts
of superoxide-related species are produced during ORR irrespective of increased rotation, and
that these species do not chemically react with DME within the timescale of transit from disk to
ring (tens of milliseconds). Such chemical stability stands in contrast to organic carbonate-based
solvents, which have been shown to readily react with 02
14'15'48
A recent RRDE study 44
calculated a first-order rate constant describing the reaction between the superoxide radical and
PC at 1 s~1, which is several orders of magnitude higher than that estimated for LiO 2
disproportionation (~ 10-1 s-1)
2,
and indicates a very high reactivity in the former case. In
contrast, the relative stability of the collection efficiency of LiO 2 indicates a relatively stable or
negligibly (< 10-3 s-1) reactive superoxide in the presence of DME.
To quantify the relative amounts of Li 2 0 2 and superoxide produced during the ORR, integrals of
the ring current during ORR (between 2.0 - 3.0 V) and disk current during OER (3.0 - 4.5 V)
with respect to time were performed to obtain the ring charge, Qring and OER charge,
respectively. By comparing these quantities against the total charge produced during ORR,
QOER
QORq,
the following relation, barring parasitic reactions with the electrolyte or other components, will
hold:
Qring + QOER
IIQORR
1
(6)
QORR
where rj represents the geometric collection efficiency of the RRDE. The first term represents the
fraction of ORR charge represented by the production of superoxide radicals, and is nearly
constant at ~ 20% during rotation as seen in Figure 12b, while the second term denotes the
fraction of Li 20 2 or other solid species such as Li 2 CO 3 on the surface of the disk, and comprises
38
the majority of the ORR charge. In the complete absence of side reactions with the electrolyte
solvent, both fractions are together expected to account for end products of the ORR and should
add up to 1.
As shown in Figure 12b, 85 - 90% of the ORR charge can be allocated to either superoxide or
the formation of Li 2 0 2 and other solid species. The remaining fraction of the ORR charge may
be attributed to incomplete removal of Li2 CO 3 or other solid species such as HCO 2 Li and
CH 3CO 2Li, which may be formed during discharge from reactions between Li 2 0 2/superoxide
radicals and carbon. 3' 19' 4 9 This result demonstrates the relative chemical stability of LiO 2 , and
provides justification for using the RRDE to track the relative contributions of LiO 2 and Li 202
production to the overall ORR charge.
3.3
Potential Dependence of Superoxide collection on Carbon
While CV measurements provide evidence of the presence of a long-lived superoxide-related
intermediate during Li-ORR and its chemical stability in DME, the continuously varying disk
potential prevents a more in-depth study of the kinetics and corresponding lifetime of the
superoxide radical during ORR on different reaction surfaces. This is particularly so since
superoxide formation on an initially pristine surface becomes coupled with superoxide formation
on electrodeposited Li 20 2 as the cathodic scan proceeds.
To systematically investigate the effect of applied discharge potential on the lifetime of the
superoxide intermediate, potentiostatic ORR measurements were performed on the disk held at
various potentials between 2.0 and 3.0 V with the ring at 3.5 V in DME, DMF and DMSO-based
electrolytes on carbon (Figure 13) and Au (Figure 15). DMSO was chosen because it is actively
39
being evaluated as an electrolyte solvent in Li-oxygen cells 26 ,50 5, 1 , and has been recently shown
to promote long cycle life (up to ~100 cycles) when used as an electrolyte solvent with a
nanoporous Au cathode.' 3 The origin of this improvement, which is not observed for analogous
Au cathodes in DME13 and DMF 39, is currently not completely understood.
Potentiostatic measurements show a clear dependence of LiO 2 stability on applied potential in
DMSO, DME and DMF. Figure 13a shows representative current responses to a step change in
disk potential from OCV to 2.6, 2.4, 2.2 and 2.0 V at a rotation rate of 900 rpm in 02 and Arsaturated DME on glassy carbon. The onset of significant ORR activity (at least 3 times above
the Ar background) begins at 2.6 V, where the disk current density is 0.041 mA/cm 2 compared to
0.0015 mA/cm2 in Ar after 12 secs (see Figure 13 inset). This is supported by the potential
dependence of Ar-corrected disk (Figure 13b) and ring (Figure 13c) charges, which likewise
display an onset at 2.6 V.
40
(a)
(b)
2.6V
-0.1
-2.4V
-21
DME
DMSO
*
-0.3 2.2V
2A-2
-0.4
-0.5
2.0V
22-V4
-0.6
4A
in
22
0
2DMF
01
2
13
4
6
10
8
GCE
4
1214
2.0
2.2
Time (ses)
2A
2.6
2.8
Potndal (Vv LVUI)
(c)
(d)
GCE
1.0
DMF
0.8
4W
~20-
25
- -
2
DMSO
OCE
-- - -
10.6
DMF
OA
DMSOI
0.2-
DME
2.0
2.2
2A
2.6
Potenial (V vs LULI')
52.8
DME
.
2.2
2A
2.8
Potential (V vs LI/Li)
2.8
Figure 13 (a) Disk current transients in Ar and 0 2 -saturated 0.1M LiClO 4 in DME at 2.6, 2.4,
2,2 and 2.0 V (b) Background-subtracted disk charge (c) Background-subtracted ring charge and
(d) Ring charge/Disk charge versus applied potential in 0.lM LiCLO 4 in DME (blue), 0.1M
LiClO 4 in DMSO (red) and 0. 1M LiClO 4 in DMF (green) on glassy carbon. RRDE rotation was
at 900 rpm.
Comparing Ar-corrected disk and ring charges among DME, DMF and DMSO, two features
stand out: (i) ring and disk charges in DME, DMF and DMSO reach maxima at 2.4, 2.5 and 2.6
V respectively; (ii) disk and ring charges at potentials greater than 2.4 V are markedly higher in
DMF than in DMSO and DME, but more comparable below 2.4 V.
Since the ring current corresponds to the oxidation of superoxide2 6 2, 7 , the first observation can be
explained by an increase in the kinetics of superoxide generation on the disk (and consequently
41
higher concentrations of superoxide oxidized on the ring) with increasing overpotential between
the ORR onset potential and potential of maximum ring/disk charge, followed by the preferential
formation of Li 20
2
on the disk at higher overpotentials. The latter regime may be dominated
either by direct electron transfer to surface superoxide or a higher rate of disproportionation from
increased local concentrations of superoxide, but will be limited to a capacity dictated by the low
conductivity of solid Li2 0 2 , as will be discussed in detail below.
The existence of predominantly superoxide and Li 20
2
formation regimes at low and high
overpotentials respectively is supported by several features of the data presented in Figure 13.
Firstly, disk charges within the superoxide regime, while too high to correspond to Li 02
2
formation, are consistent with continuous production of superoxide and its subsequent
convection to the ring. For instance, assuming that peak Ar-corrected disk charges in DME,
DMF and DMSO originate from the formation of a uniform, thin film Li 2 0 2 with density 2.36
g/cm3 results in Li 2 0 2 thicknesses of 21, 43 and 5 nm respectively. While a 5 nm Li 2O2 thickness
is a reasonable upper bound for Li 20 2 thickness, 21 and 43 nm are much higher than the 5 - 10
nm range within which conductivity limitations prevent the further deposition of Li 20 2 on planar
carbon electrodes. 35' 36 Similar assumptions for disk charges at higher overpotentials (2.0 and 2.2
V) result in more reasonable Li 20 2 thicknesses between 2.3 and 6.7 nm. It should be noted that
such thin film morphologies are only assumed for purposes of comparing product formation
across different potentials, and may not exist for these electrodes, whose morphologies are not
examined in this study.
More evidence for the predominance of soluble superoxide species at low overpotentials is seen
from oxidative linear sweep voltammetry between 3.0 and 4.5 V at 50 mV/s following
potentiostatic ORR measurements on the disk (Figure 17). The CVs show markedly lower
42
currents after potentiostatic holding at 2.7 V and above (not shown), suggesting a negligible net
deposition of Li 2 0 2 or other insoluble Li-0 2 reaction-related species during ORR at low
overpotentials. In contrast, the much higher currents observed at 2.6 V and below indicate the
electrodeposition of Li 2 0 2 or other solid species. Additionally, the ratio of Ar-corrected ring-todisk charge (Qrzng/QoRR) decreases markedly with potential (Figure 13) in DME, DMF and
DMSO. This is clearly indicative of a higher fraction of ORR charge going towards the
production of superoxide at lower overpotentials, and toward Li 2O2 or other insoluble species at
higher overpotentials.
It is particularly interesting to note that at low overpotentials (2.6 and 2.7 V) Qring/QoR in
DMSO is about 20% which begins to approach the geometric collection efficiency of the RRDE
(25%). This implies that roughly 80% (0.2/0.25) of the ORR activity is represented by
superoxide formation. In contrast, only 20% (0.05/0.25) of the ORR charge can be traced to LiO 2
at 2.0 V.
From Figure 13b, the onset of LiO 2 formation starts at 2.7 V in DMSO but at 2.6 V in DME and
DMF. This phenomenon is consistent with faster LiO 2 formation kinetics in the former case.
That this difference is fundamentally kinetic in origin is supported by CV measurements by
Laoire et a15 1 of the 02/02~ redox couple on a glassy carbon electrode in DME and DMSO where
the larger separation between anodic and cathodic peaks in DME (1.21 V) compared to DMSO
(0.23 V) was attributed to more sluggish electron transfer kinetics in the former case. In further
support of this hypothesis, galvanostatic tests in lithium-oxygen cells have shown -100 mV
higher discharge potentials in DMSO compared to ether-based electrolytes on Ketjen300 26 and
graphene-derived 5 0 carbon substrates.
43
Figure 13d shows slighter higher LiO 2 collection fractions between 2.5 and 2.7 V in DMSO and
DMF than in DME, indicating a higher chemical stability of LiO 2 in the former solvents than in
DME.
Differences in LiO 2 solubility among various electrolyte solvents have been explained in terms
of Pearson's hard soft acid base (HSAB) theory,
where hard acids are more stable against hard
bases and soft acids against soft bases. Li+, being a hard acid, would thus not be stable when
coupled to 02, which is a weak base.
In particular, Laoire and co-workers have used cyclic voltammetry to study the influence of the
electrolyte solvent donor number on the stability of the superoxide radical in the presence of Li
ions. 51 A series of CVs showing ORR and OER in Li-containing electrolytes were obtained by
limiting the cathodic scan window to successively lower potentials, and examining the resulting
anodic features in turn (Figure 14). The authors observed that in DMSO, when the cathodic scan
is limited to 2.6 V (where 2e- transfer to 02, leading to solid Li 2 0 2 formation is most limited),
there was a corresponding anodic peak (Epai) at 2.75 V (Figure 14b). The appearance of this
anodic peak below the thermodynamic potential for Li 20
2
formation (2.96 V) is clear evidence
for the stability of a superoxide-related species in DMSO, since the thermodynamic potential for
the 02/02~ couple in the absence of Li is around 2 V. 2 3
Furthermore, the absence of this corresponding anodic peak in cathodic-limited CVs in DME
(Figure 14a), acetonitrile (MeCN) and tetraglyme (TEGDME) indicates faster disproportionation
to Li 20
2
or otherwise higher superoxide reactivity. Considering that the order of solvent Lewis
basicity as characerized by the donor number is DMSO(29.8) > DME(20.0) > TEGDME(16.6) >
MeCN(14.1), the authors conclude that electrolyte solvents with high Lewis basicity are able to
44
better solvate and reduce the hard Lewis acidity of Li*, and thereby increase the lifetime the Li+02- ion
pair during oxygen reduction.
(a)
(b)
-
-
SAW
-
__VU
-
-IIIV-I-.-
*.VU'
=-
1.0152*253*354*.5
P w lid V)vL UU
ofwt
Figure 14 Cyclic voltammograms for ORR and OER in (a) 0.1M LiPF 6 in DME and (b) 0.1M
LiPF 6 in DMSO at various potential windows on glassy carbon at 100 mV/s 51
Since the collection efficiency of superoxide is directly correlated to its stability in solution, the
higher basicity of DMSO (donor number = 29.8) and DMF (donor number = 26.6) solvent
molecules compared to DME (donor number = 20.0) therefore explains higher collection
efficiencies in the former solvent. DMSO and DMF, being more electron-donating to Li+ than
DME, moderate the hard Lewis acidity of Li* ions and better stabilize the Li*-02~ pair, in a
similar manner to how TBA, which is an intrinsically weak base, is stable against 02~.
A recent RRDE study further demonstrates the ability of DMSO to better stabilize LiO 2 in
comparison to solvents with lower donor number. It was shown from CV experiments that the
addition of DMSO in a 1:200 volume ratio to an acetonitrile-based electrolyte (acetonitrile donor
number = 14.1) resulted in an increase in the collection efficiency of superoxide from 0.4% to
-25%
27
45
The existence of a potential dependence on superoxide lifetime has implications both for the
growth mechanisms and associated morphologies of Li 2 0 2 as well as parasitic reactivity during
ORR. Recent studies examining the dependence of Li 20 2 morphology on the applied discharge
rate/overpotential during ORR on vertically aligned carbon nanotubes (VACNTs)
37
and Vulcan
carbon 30 in ether-based electrolytes have found that Li 2 0 2 discs form at low applied
overpotentials or discharge rates, while thin-film deposits or coatings of Li 20
2
dominate at
higher overpotentials. It has been hypothesized on the basis of these studies that different
reaction pathways govern the growth processes of these different structures, with aggregation
and chemical disproportionation of LiO 2 species resulting in the evolution of Li 2 0 2 discs and
direct, surface-mediated 2e- transfer to 02 largely responsible for Li 2 0 2 coatings. 30
Results presented here support this hypothesis, as the collection efficiency of superoxide at very
low applied overpotentials during ORR is very close to to the geometric collection efficiency and
indicates a very high soluble superoxide stability at potentials equal to or greater than 2.6 V. This
observation correlates with surface-sensitive X-ray absorption measurements showing that Li 2 02
discs (formed at low applied overpotentials during ORR) exhibit superoxide-like/oxygen-rich
surface chemistry, while particles (formed at high applied overpotentials during ORR)
are
largely stoichiometric 37 , and is in agreement with recent DFT studies predicting that ORR at low
overpotentials is likely to result in the formation of Li 20 2 with O-rich surfaces.53
The detection of soluble superoxide species formed on carbon in DME, DMF and DMSO-based
electrolytes is in also in agreement with density functional theory (DFT) calculations suggesting
that adsorption energies of LiO 2 on the basal plane of graphitic carbon are relatively weak
(ranging from 0.29 - 0.46 eV)54 . These species are therefore likely to be highly labile during
46
ORR, and detectable by techniques such as RRDE, which rely on the convection of solutionbased species.
DFT studies have additionally suggested that, as opposed to the inert basal plane of carbon,
oxygen reduction in the presence of Li proceeds very readily on more unstable armchair or divacancy sites 54 on carbon to form Li 2 CO 3 or other products. The stability of LiO 2 in DME, DMF
and DMSO-based electrolytes during ORR on carbon may therefore explain the widespread
formation and accumulation of parasitic products such as Li 2 CO 3 , CH 3CO 2 Li and HCO 2Li in
carbon-based cathodes during lithium-oxygen cell cycling in a wide range of otherwise stable
organic3, 10 ,19, 2 0 and even ionic liquid-based55 electrolyte solvents.
3.4
Potential Dependence of Superoxide Collection on Au
On the basis of carbon reactivity, "carbon-free" cathodes based on nanoporous Au13' 39 and
mesoporous metal oxides 56' 57 are actively being considered for Li-Air batteries. Since the
adsorption energy/reactivity of LiO 2 will vary depending on the nature of the reaction
surface58 ' 59 , an examination of LiO 2 formation kinetics and lifetimes on these various proposed
potential cathode surfaces is crucial to assessing their feasibility for practical applications.
The potential dependence of LiO 2 generation on Au in DME and DMSO exhibits some
differences from the analogous case on glassy carbon. Figure 15a shows representative current
responses to a step change in disk potential from OCV to 2.6, 2.4, 2.2 and 2.0 V at a rotation rate
of 900 rpm in 02 and Ar-saturated DME on Au. Similar to results obtained on glassy carbon,
disk and ring charge maxima are observed, as well as higher overall ORR activity in DMSO and
DMF relative to DME. The onset of ORR activity in DME begins at 2.5 V, while the analogous
47
process on glassy carbon starts at 2.6 V. Similarly, the peak ORR disk charge on Au is 0.56 mC
at 2.2 V, in contrast to that of glassy carbon at 0.92 mC at 2.4 V, suggesting that the kinetics of
LiO 2 formation are slower on Au in DME than on glassy carbon in DME.
(b)
(a)
-0.1 2.4V
2.2V
E
-0.22
'IV
LM
-.0.3
4
2.OV
U1
12
13
-0.4
0
2
4
6
8
10
12
Potential (V vs LI/LI)
Potential (V vs LWLI')
(d)
0
0I..U
U
I*1
S
2.0
2.2
2.4
2.6
Potential (V vs LI')
2.8
2.0
2.2
2.4
2.6
2.8
Potential (V vs LULI')
Figure 15 (a) Disk current transients in Ar and 0 2 -saturated 0.1M LiClO 4 in DME at 2.6, 2.4,
2,2 and 2.0 V (b) Background-subtracted disk charge (c) Background-subtracted ring charge and
(d) Ring charge/Disk charge versus applied potential in 0.1M LiClO4 in DME (blue) and 0.1M
LiClO 4 in DMSO (red) on Au at 900rpm.
The origin of this difference is currently unknown, but is likely related to one of the following
possibilities. The first is the tendency for metal electrodes to interact strongly with the
superoxide radical, or otherwise render the 02/02 redox couple more irreversible than it would
48
be on carbon as shown by Sawyer et al2 3 and in Figure 4. Following HSAB theory, since DME is
a weak Lewis base, it will poorly solvate Li+ ions, and lead to a more unstable Lie-02~ pair with
higher reactivity with Au during ORR. Another explanation could be the susceptibility of
polycrystalline Au to form complex passivating films during electrochemical cycling in DME.14
This latter possibility is particularly relevant to this study as the Au electrode is
electrochemically "cleaned" of any deposited material by cycling between 3.0 and 4.0 V after
each potentiostatic ORR measurement in 0 2 -saturated electrolyte. Au hydroxide has been shown
by Fourier-Transform Infrared Spectroscopy (FTIR) to form above 3.2 V 14 , and might therefore
subtly influence the kinetics of LiO 2 formation over time.
While the current response and associated ring and disk charges during ORR on Au in DMF are
similar to that of glassy carbon in DMF, similar measurements in DMSO show significant
differences between glassy carbon in DMSO. Both surfaces exhibit peak ring and disk charges at
2.6 V, but no net generation of superoxide below 2.5 V in the case of Au (Figure 15).
49
~25
- - - - - --------- --Carbon
20
15
5 -Au
2.0 2.1 2.2 2.3 2.4 2.5 2.6 2.7 2.8 2.9
Potential (V vs LIILI)
Figure 16 Ring charge/Disk charge versus applied potential in 0.1 M LiClO 4 in DMSO on carbon
(filled circles) and Au (open circles) at 900rpm. Error bars are calculated from 3 independent
measurements.
This is indicative of a more kinetically facile pathway to the formation of Li 2 0 2 on Au in DMSO
than exists on Au in DME or on carbon. A recent investigation by Peng et al 3 of Li-0
2
cell
electrochemistry using a nanoporous Au cathode showed by FTIR and Raman spectroscopy that
Li 2 O2 is the primary discharge product during continuous cycling in 0.1 M LiClO 4 in DMSO.
50
2.4 Vso
100
16
102.
40
0
20 t2.7
2.4 V
j20
2.6
0
.
V
00
3.2
3A3.6
3.8
4. 4.44.6
SA
3.2
Potenial (Vvs UIU*)
140 (c)
2.0 Vso
1
V02.
100 1002.
so
3.6
Potendal (Vvs
2.6 V
(d)
o
3.8
4.0
LU*)
2.0 V
.
48
2.6 V
20
-202.V2.V
3.2
3A
3.6 3.8 4.0 4.2
PotSndal (V vs LIllX)
4.4
4.6
3.2
3.6
3.4
3.6
Potndal (V vs UL&*)
4.0
Figure 17 Linear sweep voltammetry measurements at 50 mV/s following potentiostatic ORR at
2.7, 2.6, 2.4 and 2.0 V on (a) glassy carbon in 0.1M LiClO 4 in DME (b) Au in 0.1M LiClO 4 in
DME (c) glassy carbon in 0.1M LiClO 4 in DMSO (d) Au in 0.1M LiClO 4 in DMSO to oxidize
electrodeposited species.
Additionally, in contrast to DMSO, they also showed that ORR on an Au electrode in 0.1M
LiClO 4 in DME led to significant electrolyte decomposition1.
This observation is consistent
with the results in Figure 15d, showing that LiO 2 generated on Au in DME is more stable/longlived than that formed in DMSO, and is therefore more prone to electrolyte attack over the
typical timescale of galvanostatic discharge.
51
GCE
(a)
(D)
25
e25 -----------------20
20
DMSO
DMF
15.
Au
DME
-15
10
10.
cc0
011
2.0
DMSO
DMF
5
DME
5
2.2
2.4
2.6
Potential (V vs LULI)
2.8
2.0
2.2
2.4
2.6
2.8
Potential (V vs LLl')
Figure 18 Separate measurements of ring charge/disk charge versus applied potential in 0.1M
LiClO 4 in DMSO, DMF and DME on (a) glassy carbon and (b) Au
52
3.5
Mechanistic Implications
The stability of the superoxide radical in the three electrolyte solvents studied may be explained
by accounting for two sets of interactions: between the superoxide radical and the electrolyte
solvent and superoxide and electrode surface. In the first case, the lifetime of the superoxide
radical is directly related to the ability for solvent molecules to acts as electron donors in order to
moderate the hard Lewis acidity of Li+ and stabilize the Li-02 bond, in accordance with HSAB
theory. Such stability thereby reduces, but does not completely eliminate, the propensity for LiO 2
to attack electrolyte solvents through nucleophilic reactions. 10' 14'39 In the case of electrode
interactions, the adsorption energy/degree of interaction between superoxide and the reaction
surface is of critical importance. Higher adsorption energies favor the stabilization of superoxide
on the surface of the electrode, where further reduction or disproportionation to Li 20 2 may occur
during ORR.
Since the adsorption energy of LiO 2 on the basal plane of carbon is very low (0.29 - 0.5 eV) 54 ,
superoxide radicals produced on carbon are highly labile, and better able to react with and
decompose a wide range of electrolyte solvents, including those evaluated in this study. Even in
the case of electrolytes considered highly stable, such as ionic-liquid based configurations55
superoxide radicals can react with the more unstable oxidized COx sites 54 on carbon to form side
products that may then accumulate during cycling.
53
This explains the general similarity in the potential dependence of superoxide stability during
ORR in DMSO, DMF and DME on glassy carbon, where the lifetime of superoxide is
determined not by solvent properties, but by the applied ORR potential: at low overpotentials,
the kinetics of Li 20 2 formation from LiO 2 are slow and therefore a higher fraction of total ORR
charge is detected in the superoxide form. These differences in stability provide support for two
different growth mechanisms responsible for toroidal and thin film morphologies of Li 2 0 2 , as
discussed above. The high mobility of LiO 2 in this scenario may also explain the widespread
evidence of accumulation of side products and associated capacity fade observed in the vast
0 15 16
majority of carbon-based Li-0 2 cathodes reported in the literature. , , ,20
In contrast to carbon, superoxide stability on Au exhibits a strong dependence on the nature of
the electrolyte solvent. DFT studies 58' 59 and cyclic voltammetry2 3 experiments confirm that the
adsorption energy/interaction of superoxide on Au is higher than that of carbon. Thus,
superoxide produced on Au will have a greater tendency to interact with the electrode surface
and be reduced to Li 2 02, as opposed to being solubilized by the electrolyte, and promoting side
reactions.
This observation, when coupled with the effect of the Lewis basicity of electrolyte solvent
basicity (represented by donor number) on the stability of Li*-02~ ion pair, explains the
superoxide stability trend obtained on Au in DMSO, DMF and DME. Because DMSO has the
highest donor number (29.8) of the solvents studied, it is best able to stabilize the Li*-02~ ion
pair, and thereby allow maximal interaction between the superoxide ion and the surface of Au,
leading to its most efficient reduction/disproportionation to Li 20 2 , and hence little detection of
LiO 2 below 2.5 V. DME, which has the lowest donor number of 20.0 is least able to stabilize the
54
Li*-02~ ion pair and therefore permits the highest fraction of LiO 2 to be detected in solution even
at high overpotentials. DMF, with a donor number of 26.6, represents the intermediate case.
Understanding the effect of the nature of the reaction surface and electrolyte solvents on the
stability of LiO 2 may shed light on reaction mechanisms in practical lithium-oxygen cells.
However, systematic studies of Li-0
2
reaction intermediate stability using spectroscopic
techniques represent one route to extending electrochemical insights from model systems to real
cells. These will be discussed in detail below.
55
56
4 Conclusions and Future Work
Fundamentally understanding reaction mechanisms during Li-0
2
cell operation is critical for
implementing battery chemistries with high reversibility and long cycle life. Central to this task
is investigating the lifetime of Li-0 2 reaction intermediates, since they are thought to contribute
to the poor chemical stability of cell components such as the cathode surface, electrolyte and
binder. In this thesis, the RRDE technique has been used to probe the influence of electrolyte
solvent and reaction surface on superoxide stability.
It was found that the relative quantity of the superoxide radical generated during ORR exhibits
potential dependence, such that more superoxide is produced as a fraction of total ORR charge at
low overpotentials than at high overpotentials. This basic trend is observed in DMSO, DMF and
DME-based electrolytes, and indicates that the lifetime of the superoxide radical is higher at
lower overpotentials than at higher overpotentials. Understanding the lifetime of superoxide at
different applied potentials is particularly relevant to recent experimental studies suggesting that
the growth processes of Li 2 0
2
discs formed at low overpotentials are governed by the
aggregation and disproportionation of LiO 2 , while thin deposits/small particles of Li 2 0 2 originate
chiefly from surface-mediated electron transfer to LiO 2 .30 The higher lifetime of LiO 2 at lower
overpotentials is consistent with this growth mechanism, and also provides a path to explaining
why Li 2 O2 discs have 0-rich/superoxide-like surface chemistry, and particles are largely
stoichiometric. 37
Several important differences were observed upon ORR in DMSO, DMF and DME electrolytes.
On both carbon and Au, superoxide formation was found to begin at higher potentials in DMSO
than in DMF and DME, indicating that LiO 2 formation has the fastest kinetics in DMSO. While
57
this result is consistent with previously reported observations 26,50 that discharge voltage plateaus
are ~100 mV higher in DMSO than in ether-based electrolytes, more systematic studies
explicitly relating superoxide formation to applied potential in practical cells are necessary to
validate the hypothesis that faster superoxide formation kinetics are responsible for this
difference.
In addition to the role of the solvent, the electrode surface was also found to influence the ORR
kinetics and pathways. It was found that no LiO 2 was produced below 2.5 V on an Au surface,
and that the kinetics of Li 2 0 2 formation was much faster on Au in DMSO than on Au in DME.
This difference may be explained in terms of the effect of solvent basicity on the stability of
LiO 2 and suggests higher stability of superoxide on Au in DMSO in comparison to Au in DME
in DMF. This is consistent with a higher degree of electrolyte attack in DME and DMF, and is in
agreement with recent experimental studies showing that Li-0
2
electrochemistry is more
reversible in DMSO than in DMF 39 and DME.1 3 This highlights the important relationship
among LiO 2 lifetime, the formation of parasitic species and the ultimate reversibility of Li-0
2
electrochemistry for a given set of cell parameters/chemistries.
The work in this thesis demonstrates the use of the RRDE both as a method for probing Li-0
2
reaction intermediate reactivity under different conditions, and as screening tool for predicting
Li-0
2
cell configurations most prone to the formation of parasitic species during ORR, and hence
progressively higher OER overpotentials during cycling. 19
Future work should also include studies that couple electrochemical with in situ spectroscopic
investigation of the Li-0 2 reaction, where reaction intermediates and products can be directly
identified. A recent in situ ambient pressure X-ray Photoelectron Spectroscopy (APXPS) study
58
of the Li-0 2 reaction in a solid-state cell6 o represents a step in this direction. In this investigation,
reversible formation and decomposition of Li 20 2 was visualized for the first time, in the absence
of potentially reactive electrolyte and binder components. Applying this concept to techniques
that are more sensitive than APXPS to element-specific electronic and local structure such as Xray Absorption Near Edge Structure (XANES) will give valuable information on the intrinsic
relationship between surface and bulk chemistries of the reaction product and associated ORR
and OER overpotentials.
In order to in probe reaction mechanisms under more practical conditions, i.e. including
electrolyte solvent, vibrational methods such as Raman and FTIR are more appropriate. Raman
spectroscopy has been successfully used for in situ characterization of reaction intermediates and
products, but has limited sensitivity in detecting species in the presence of carbon, whose
prominent D and G vibrational bands make the detection of other species difficult. 61
Spectroscopic methods such as electron paramagnetic resonance (EPR) may prove useful in
better identifying reaction intermediates, especially in solution. EPR is particularly suited to the
detection of paramagnetic species such as the superoxide radical 4 ' both as a solution-based
molecular species, and in solid-state form. 62 In situ electrochemical studies involving EPR would
therefore be able to more explicitly couple the ORR overpotential dependence of various
observed Li 20 2 morphologies to different lifetimes of LiO 2 . Additionally EPR studies during
OER would grant the ability to infer the evolution of singlet oxygen, which has been speculated
to promote electrolyte degradation 63 during radiative transitions to ground state, EPR-active
triplet oxygen.
59
60
References
Tarascon, J.-M.; Armand, M. Nature 2001, 414, 359.
(1)
Wagner, F. T.; Lakshmanan, B.; Mathias, M. F. 2010.
(2)
Ottakam Thotiyl, M. M.; Freunberger, S. A.; Peng, Z.; Bruce, P. G. Journal of the
(3)
American Chemical Society 2012, 121128114715003.
Viswanathan, V.; Norskov, J. K.; Speidel, A.; Scheffler, R.; Gowda, S.; Luntz, A.
(4)
C. The Journalof Physical Chemistry Letters 2013, 556.
Chen, H.; Armand, M.; Demailly, G.; Dolhem, F.; Poizot, P.; Tarascon, J.-M.
(5)
ChemSusChem 2008, 1,348.
Kang, K. S.; Meng, Y. S.; Breger, J.; Grey, C. P.; Ceder, G. Science 2006, 311,
(6)
977.
Wang, Z.-L.; Xu, D.; Xu, J.-J.; Zhang, L.-L.; Zhang, X.-B. Adv. Funct.Mater.
(7)
2012, 22, 3699.
Mitchell, R. R.; Gallant, B. M.; Thompson, C. V.; Shao-Horn, Y. Energy &
(8)
Environmental Science 2011, 4, 2952.
Lu, Y.-C.; Kwabi, D. G.; Yao, K. P. C.; Harding, J. R.; Zhou, J.; Zuin, L.; Shao(9)
Horn, Y. Energy & EnvironmentalScience 2011, 2999.
Freunberger, S. A.; Chen, Y.; Drewett, N. E.; Hardwick, L. J.; Barde, F.; Bruce,
(10)
P. G. Angew Chem Int Edit 2011, 50, 8609.
Lee, J.-H.; Black, R.; Popov, G.; Pomerantseva, E.; Nan, F.; Botton, G. A.; Nazar,
(11)
L. F. Energy & EnvironmentalScience 2012, 5, 9558.
Younesi, R.; Hahlin, M.; Treskow, M.; Scheers, J.; Johansson, P.; Edstrom, K.
(12)
The Journalof Physical Chemistry C 2012, 120807123830006.
Peng, Z.; Freunberger, S. a.; Chen, Y.; Bruce, P. G. Science (New York, N.Y)
(13)
563.
2012, 337,
Aurbach, D.; Daroux, M.; Faguy, P.; Yeager, E. JournalofElectroanalytical
(14)
Chemistry and InterfacialElectrochemistry 1991, 297, 225.
Freunberger, S. A.; Chen, Y. H.; Peng, Z. Q.; Griffin, J. M.; Hardwick, L. J.;
(15)
Barde, F.; Novak, P.; Bruce, P. G. Journalof the American Chemical Society 2011, 133, 8040.
Mccloskey, B. D.; Bethune, D. S.; Shelby, R. M.; Girishkumar, G.; Luntz, A. C. J
(16)
Phys Chem Lett 2011, 2, 1161.
Bryantsev, V.; Faglioni, F. JPCA 2012, 116.
(17)
(18)
Zhang, Z.; Lu, J.; Assary, R. S.; Du, P.; Wang, H.-H.; Sun, Y.-K.; Qin, Y.; Lau,
K. C.; Greeley, J.; Redfern, P. C.; Iddir, H.; Curtiss, L. A.; Amine, K. Physical Chemistry
Chemical Physics 2011, 15.
Gallant, B. M.; Mitchell, R. R.; Kwabi, D. G.; Zhou, J.; Zuin, L.; Thompson, C.
(19)
V.; Shao-Horn, Y. The Journalof Physical Chemistry C 2012, 116, 20800.
McCloskey, B. D.; Speidel, A.; Scheffler, R.; Miller, D. C.; Viswanathan, V.;
(20)
Hummelshoj, J. S.; Norskov, J. K.; Luntz, A. C. The JournalofPhysical Chemistry Letters 2012,
997.
Black, R.; Oh, S. H.; Lee, J. H.; Yim, T.; Adams, B.; Nazar, L. F. Journalof the
(21)
American Chemical Society 2012, 134, 2902.
Younesi, R.; Hahlin, M.; Bj6refors, F.; Johansson, P.; Edstr6m, K. Chemistry of
(22)
Materials2012, 77.
61
(23)
Sawyer, D. T.; Chiericato, G.; Angells, C. T.; Nannl, E. J.; Tsuchlya, T.
Analytical Chemistry 1982, 2, 1720.
(24) Lu, Y.-C.; Gallant, B. M.; Kwabi, D. G.; Harding, J. R.; Mitchell, R. R.;
Whittingham, M. S.; Shao-Horn, Y. Energy & Environmental Science 2013, 6, 750.
(25) Peng, Z.; Freunberger, S. A.; Hardwick, L. J.; Chen, Y.; Giordani, V.; Bard6, F.;
Novik, P.; Graham, D.; Tarascon, J.-M.; Bruce, P. G. Angewandte Chemie (Internationaled in
English) 2011, 50, 6351.
(26) Trahan, M. J.; Mukerjee, S.; Plichta, E. J.; Hendrickson, M. A.; Abraham, K. M.
Journalof the ElectrochemicalSociety 2012, 160, A259.
(27) Calvo, E. J.; Mozhzhukhina, N. Electrochemistry Communications 2013, 31.
(28) McCloskey, B. D.; Scheffler, R.; Speidel, A.; Girishkumar, G.; Luntz, A. C. J
Phys Chem C 2012, 116, 23897.
(29) Fan, W.; Cui, Z.; Guo, X. The JournalofPhysicalChemistry C 2013,
130130163402009.
(30) Nazar, L.; Adams, B.; Black, R.; Radtke, C.; Zaghib, K.; Trudeau, M. Energy &
EnvironmentalScience 2013.
(31)
Mitchell, R. R.; Gallant, B. M.; Shao-Horn, Y.; Thompson, C. V. Journalof
PhysicalChemistry Letters 2013, 4, 1060.
(32)
Radin, M. D.; Rodriguez, J. F.; Tian, F.; Siegel, D. J. Journalof the American
Chemical Society 2012, 134, 1093.
(33)
Ong, S.; Mo, Y.; Ceder, G. Physical Review B 2012, 85, 2.
(34) Hummelshoj, J. S.; Blomqvist, J.; Datta, S.; Vegge, T.; Rossmeisl, J.; Thygesen,
K. S.; Luntz, a. C.; Jacobsen, K. W.; Norskov, J. K. The Journalof chemical physics 2010, 132,
071101.
(35)
Viswanathan, V.; Thygesen, K. S.; Hummelshoj, J. S.; Norskov, J. K.;
Girishkumar, G.; McCloskey, B. D.; Luntz, A. C. J Chem Phys 2011, 135, 214704.
(36) Albertus, P.; Girishkumar, G.; McCloskey, B.; Sanchez-Carrera, R. S.; Kozinsky,
B.; Christensen, J.; Luntz, A. C. Journalof the ElectrochemicalSociety 2011, 158, A343.
(37)
Gallant, B. M.; Kwabi, D. G.; Mitchell, R. R.; Zhou, J.; Thompson, C. V.; ShaoHorn, Y. Submitted 2013.
(38)
Read, J. Journalof The ElectrochemicalSociety 2002, 149, Al 190.
(39)
Chen, Y.; Freunberger, S. A.; Peng, Z.; Barde, F.; Bruce, P. G. Journalof the
American Chemical Society 2012.
(40) Read, J.; Mutolo, K.; Ervin, M.; Behl, W.; Wolfenstine, J.; Driedger, A.; Foster,
D. Journalof The ElectrochemicalSociety 2003, 150.
(41)
Maricle, D. L.; Hodgson, W. G. Analytical Chemistry 2002, 37.
(42)
Sawyer, D. T.; Valentine, J. S. Accounts of Chemical Research 2002, 14.
(43)
Bard, A. J.; Faulkner, L. R. ElectrochemicalMethods: Fundamentalsand
Applications; Wiley, 2004.
(44) Herranz, J.; Garsuch, A.; Gasteiger, H. A. The JournalofPhysical Chemistry C
2012, 120807071402004.
(45) Paulus, U. A.; Schmidt, T. J.; Gasteiger, H. A.; Behm, R. J. Journalof
ElectroanalyticalChemistry 2001, 495, 134.
(46) Viswanathan, V.; Thygesen, K. S.; Hummelshoj, J. S.; Norskov, J. K.;
Girishkumar, G.; McCloskey, B. D.; Luntz, a. C. The Journalof chemicalphysics 2011, 135,
214704.
62
(47)
Lu, Y.-C.; Gasteiger, H. A.; Crumlin, E.; Robert McGuire, J.; Shao-Horn, Y.
Journalof The ElectrochemicalSociety 2010, 157.
(48)
McCloskey, B. D.; Bethune, D. S.; Shelby, R. M.; Girishkumar, G.; Luntz, A. C. J
Phys Chem Lett 2011, 2, 1161.
McCloskey, B. D.; Speidel, A.; Scheffler, R.; Miller, D. C.; Viswanathan, V.;
(49)
Hummelshoj, J. S.; Norskov, J. K.; Luntz, A. C. JPhys Chem Lett 2012, 3, 997.
(50)
Xu, D.; Wang, Z.-l.; Xu, J.-j.; Zhang, L.-l.; Zhang, X.-b. Chem. Commun. 2012,
48, 6948.
(51)
Laoire, C. 0.; Mukerjee, S.; Abraham, K. M.; Plichta, E. J.; Hendrickson, M. A.
The Journal of Physical Chemistry C 2010, 114, 9178.
(52)
(53)
Pearson, R. G. Journalof the American Chemical Society 2002, 85, 3533.
Hummelshoj, J. S.; Luntz, A. C.; NoI'skov, J. K. The Journalof chemicalphysics
2013, 138, 034703.
(54)
Xu, Y.; Shelton, W. A. Journal of The ElectrochemicalSociety 2011, 158, Al 177.
(55)
Mizuno, F.; Takechi, K.; Higashi, S.; Shiga, T.; Shiotsuki, T.; Takazawa, N.;
Sakurabayashi, Y.; Okazaki, S.; Nitta, I.; Kodama, T.; Nakamoto, H.; Nishikoori, H.; Nakanishi,
S.; Kotani, Y.; Iba, H. JournalofPower Sources 2013, 228, 47.
(56)
Oh, S. H.; Nazar, L. F. A dvanced Energy Materials 2012, 2, 903.
(57)
Riaz, A.; Jung, K.-N.; Chang, W.; Lee, S.-B.; Lim, T.-H.; Park, S.-J.; Song, R.-H.;
Yoon, S.; Shin, K.-H.; Lee, J.-W. Chemical communications 2013.
(58)
(59)
Dathar, G. K. P.; Shelton, W. A.; Xu, Y. JPhys Chem Lett 2012,3, 891.
Xu, Y.; Shelton, W. A. J Chem Phys 2010, 133.
(60)
Lu, Y.-C.; Crumlin, E. J.; Veith, G. M.; Harding, J. R.; Mutoro, E.; Baggetto, L.;
Dudney, N. J.; Liu, Z.; Shao-Horn, Y. Sci. Rep. 2012, 2.
Yang, J.; Zhai, D.; Wang, H.-H.; Lau, K. C.; Schlueter, J. A.; Du, P.; Myers, D. J.;
(61)
Sun, Y.-K.; Curtiss, L. A.; Amine, K. Physical Chemistry Chemical Physics 2013, 15.
Gerbig, 0.; Merkle, R.; Maier, J. Advanced materials (DeerfieldBeach, Fla.)
(62)
2013.
Hassoun, J.; Croce, F.; Armand, M.; Scrosati, B. Angewandte Chemie
(63)
(Internationaled. in English) 2011, 50, 2999.
63
Download