Sie sind auf Seite 1von 30

Review

pubs.acs.org/CR

Aprotic and Aqueous LiO2 Batteries


Jun Lu, Li Li, Jin-Bum Park, Yang-Kook Sun,*, Feng Wu,*, and Khalil Amine*,,

Chemical Sciences and Engineering Division, Argonne National Laboratory, 9700 South Cass Avenue, Lemont, Illinois 60439,
United States

Beijing Key Laboratory of Environmental Science and Engineering, School of Chemical Engineering and the Environment, Beijing
Institute of Technology, Beijing 100081, China

Department of Energy Engineering, Hanyang University, Seoul 133-791, South Korea

Chemistry Department, Faculty of Science, King Abdulaziz University, 80203 Jeddah, Saudi Arabia
1. INTRODUCTION
Currently, fossil fuels supply over 85% of the worlds evergrowing energy demand.1 There is an increasing concern about
the global climate change resulting from the worldwide use of
fossil fuels, which release large quantities of CO2 and other
greenhouse gas (GHG) to the atmosphere.1 The petroleum
that is used for automobile and light truck applications
represents 34% of the worlds total primary energy source. In
the United States, the transportation sector is the single greatest
consumer of imported oil. In 2010, 94% of U.S. transportation
energy was derived from petroleum, nearly half of which came
from foreign sources.2 In terms of economic impact, petroleum
imports represented nearly 41% of the $646 billion U.S. trade
decit in 2010.3 The CO2 emissions due to the U.S.
transportation sector account for 40% of the total CO2
emission, which is considered as a major cause of geopolitical
instability.4 The U.S. transportation sector also represents
about 27% of all U.S. GHG emissions.5 However, even with
todays mix of fossil, nuclear, and renewable energy sources for
U.S. electric power generation, it is estimated that, on a well-towheel basis, an all-electric vehicle will generate 25% less GHG
emissions than a conventional gasoline-powered vehicle.6 Even
lower emissions are predicted with increased use of renewable
energy sources. Therefore, it would greatly benet the United
States and the world to transition to an electried transportation system, which is already beginning with the advent of
hybrid electric vehicles (HEVs) and will accelerate as plug-in
hybrid electric vehicles (PHEVs) and ultimately pure electric
vehicles (PEVs) gain a larger share of the market. In addition,
PEVs have the promise to greatly improve energy eciency.
For example, on a well-to-wheel basis, the all-electric Tesla
Roaster charged with electricity generated from natural gas has
an eciency of 1.14 km/MJ, nearly 2 times as ecient as a
Toyota Prius hybrid (0.56 km/MJ) and 4 times more ecient
than a typical gasoline-powered car, such as the Toyota Camry
(0.28 km/MJ).7 The use of alternative energy sources, such as
nuclear, solar, and wind power, would reduce our dependence
on fossil fuels and, thus, also reduce CO2 emissions, but devices
to store the electric energy generated by these power plants are
sorely needed. One of the most viable candidates for such
devices is the rechargeable Li battery.
Electrical energy storage is attracting signicantly more
interest nowadays, considering the expanding market for

CONTENTS
1. Introduction
2. Overview
3. Aprotic LiO2 Battery
3.1. Electrochemical Reactions in the Aprotic Li
O2 System
3.2. Early-Stage Research on the Aprotic LiO2
Battery
3.3. Eect of the Electrolyte
3.3.1. Lessons Learned from Organic Carbonate-Based Electrolytes
3.3.2. Ether-Based Electrolyte
3.3.3. Other Organic Electrolytes
3.3.4. Eect of Lithium Salt
3.3.5. HardSoft AcidBase (HSAB) Theory
3.3.6. Brief Summary
3.4. Electronic and Magnetic Properties of Li2O2
and Relevance to the LiO2 Battery
3.5. Air Electrode Architecture
3.5.1. Porous Carbon
3.5.2. Catalysts on Porous Carbon
3.6. Lithium Electrode
4. Aqueous LiO2 Batteries
5. Concluding Remarks and Perspectives
Author Information
Corresponding Authors
Notes
Biographies
Acknowledgments
References

A
B
D
D
E
F
G
H
K
M
M
N
N
O
P
R
T
V
W
X
X
X
Y
Z
Z

Received: October 16, 2013

XXXX American Chemical Society

dx.doi.org/10.1021/cr400573b | Chem. Rev. XXXX, XXX, XXXXXX

Chemical Reviews

Review

We start by presenting an overview of the LiO2 battery in


section 2, including operation principles, dierent cell
chemistries, and primary development challenges. The main
focus of section 3 is the aprotic LiO2 system, since it has been
demonstrated to possess the greatest potential to meet the PEV
requirements and has dominated the Liair battery research
eorts in the past decade. This section discusses electrochemical reactions in the aprotic LiO2 system, the stability of
aprotic electrolytes and its eect on cell performance, the
electronic and magnetic properties of Li2O2 and their relevance
to the aprotic LiO2 battery, the importance of the O2breathing electrode, and the eect of O2 crossover on the
stability of the lithium electrode. In section 4, we briey discuss
aqueous Liair systems and the challenges they are facing.
Section 5 presents concluding remarks and prospects for the
future development of Liair batteries. Covering the immense
body of all the work published in this eld is, however, beyond
the scope of this review.

portable electronics, the electrication of the transportation


sector as a result of the commercialization of battery-powered
vehicles, and the greater use of batteries for grid energy
storage.8 The specic energy (energy per unit mass) and energy
density (energy per unit volume) for state-of-the-art Li ion
batteries are much too low for all-electric vehicles, even with
the assumption that the theoretical capacity of the electrode
materials can be achieved.9 Given todays automobile Li ion
cells, the driving range is limited to about 70 miles for a 200 kg
battery pack, assuming a specic energy of 150 Wh/kg at the
cell level and 105 Wh/kg at the pack level (70% of a packs
mass is due to the cells) can be realized. With the recent
intensive development of high-capacity positive electrode
materials (275 mAh/g) and high-capacity alloy negative
electrode materials (2000 mAh/g), the specic energy of the
Li ion batteries could be eventually pushed up to 400 Wh/kg.
However, even with such high capacity, the vehicle driving
range is only doubled at the most (140 miles). The U.S.
Advanced Battery Consortium (USABC) set the goals for EVs
at a calendar life of 15 years and operating temperature from
30 to +52 C with a driving range of 300 miles per single
charge, which is beyond the electrochemical performance of
todays lithium ion batteries. To produce electric vehicles with a
range comparable to that of todays vehicles powered by liquid
fuels, a battery system that has much higher specic energy and
energy density is required. Therefore, an increasing amount of
recent research has been devoted to energy storage systems that
can go beyond the Li ion battery limits.9
One such technology is the Liair battery, which is based on
the LiO2 electrochemical couple.10 A lithiumair cell when
discharged to the lithium peroxide composition (Li2O2) at an
average potential of 3.1 V would provide a theoretical specic
energy of 3623 Wh/kg and when discharged to Li2O at the
same potential would provide a theoretical specic energy of
5204 Wh/kg. Originally proposed in the 1970s as a possible
power source for electric vehicles,1121 Liair batteries
recaptured signicant scientic interest in the late 2000s due
to advances in materials technology and an increasing demand
for environmentally safe and oil-independent energy sources.
Interest has increased sharply recently, as evidenced by over
300 research papers having been published on this topic in the
past 3 years alone. This intense research activity can be
attributed to the high energy density of the Liair battery,
potentially up to 23 kWh/kg on the cell level, and the open
cell conguration that uses air as the reactant. A fully developed
Liair battery system is expected to truly surpass the current
battery technology, even that under development for deployment in the medium term (400 Wh/kg), and meet the
requirements for the PEV application.
Development of a practical Liair battery will involve
overcoming many formidable challenges, including the need for
a fundamental understanding of LiO2 electrochemistry,
development of new and improved cell materials, and
innovation in the critical aspects of cell design. In the past
few years, dozens of reviews on the topic of Liair
batteries9,2243 have been published. These reviews address
the technical issues and challenges facing Liair batteries at the
current stage from dierent perspectives, including the stability
of the electrolytes, importance of the air electrode/electrocatalyst, and oxygen-selective membranes. In this review, we
mainly focus on the most critical issues that must be addressed,
with the hope that it will help to advance a truly rechargeable
Liair battery toward its practical application.

2. OVERVIEW
Before presenting the main topic of this review, we need to
clarify some terminology. First, it is not precise to refer to an
aprotic solvent as nonaqueous in most of the relevant
literature because, strictly speaking, nonaqueous solvents
include aprotic and protic solvents. The term aprotic is
used throughout this review since almost all the electrolytes
investigated so far are based on the aprotic solvent. We do not
consider protic nonaqueous Liair battery systems because of a
lack of published work on this topic. Likewise, mixtures of
water and other protic solvents are beyond the scope of this
review. From an electrochemical perspective, the protic system
is expected to share general characteristics with the aqueous
Liair chemistry. Also, the term Liair battery with the
aprotic system has been widely accepted and used by many
other researchers but does not precisely represent what
happens in these cells at the current time, considering that
most laboratory work has been performed under a pure oxygen
environment. This is because other components in air such as
H2O and CO2 could interfere with the desired electrochemical
behavior and, therefore, degrade the overall performance of the
battery. In light of the above concern, we refer to LiO2
battery throughout this review when we discuss the aprotic
Liair system. This distinction has been recognized in some
recently published papers.4447 However, it should be
emphasized that a true Liair battery is still the ultimate
goal,48,49 as long as a selective membrane can be developed to
prevent the permeation of other gases from the air rather than
oxygen gas. In addition, the expression of the specic capacity
of the LiO2 battery is specically claried in this review. The
term mAh/g is commonly used in most of the relevant
literature, some based on the mass of active materials and
others based on the mass of the air electrode support. The
value based on the electrode support mass is reliable only when
the capacity is shown to be proportional to the support mass. In
this sense, the value based on the active materials is more
appropriate to describe the specic capacity of the LiO2 cell at
the current moment. Alternatively, the capacity per surface area
(mAh/cm2) of the support is also accepted to describe the
specic capacity of the cell. However, the value based on the
surface area of the electrode could possibly induce improper
conclusions if the loading of the active materials is signicantly
dierent from case to case. In this paper, we adopt the term
mAh/g based on the active materials to describe the specic
B

dx.doi.org/10.1021/cr400573b | Chem. Rev. XXXX, XXX, XXXXXX

Chemical Reviews

Review

capacity of the cell. Likewise, the rate expression mA/g based


on the active materials is used in this paper for the same reason.
Practically, the energy density on the cell level would be a more
reliable value to demonstrate this promising technology.
The LiO2 battery chemistry uses the oxidation of lithium at
the metal electrode and reduction of oxygen at the air electrode
to induce current ow. The major appeal of the LiO2 battery
is its extremely high energy density, which rivals that of
traditional gasoline-powered engines. This advantage in energy
density is derived by utilizing the oxygen in air, eliminating the
need to store reactant (oxygen) at the air electrode. On the
basis of the oxidation of 1 kg of lithium metal, the theoretical
energy density of a LiO2 cell is calculated to be 11 680 Wh/
kg, which is not much lower than that of gasoline (13 000 Wh/
kg), as shown in Figure 1.35 Practically, the achievable energy

aprotic54 (see Figure 2). For all types of LiO2 systems, an


open system is required to obtain oxygen from the air because
oxygen is the active material of the air electrodes. Li metal must
also be used as the metal electrode to provide the lithium
source for all the systems at the current stage. In an aprotic Li
O2 cell, porous carbon must be added as the reservoir for the
insoluble discharge products, presumably Li2O2. Most of the
time, electrocatalysts are essential to promote the oxygen
reduction and oxygen evolution reactions during the cell
discharge and charge. In the case of the aqueous and hybrid
aqueous/aprotic systems, a protective layer for Li metal, which
prevents the vigorous reaction of lithium with water, is
necessary to enable the desired electrochemistry. The chemistry
at the oxygen electrode diers depending on the electrolyte.
Aqueous and hybrid systems share the same reaction
mechanisms since the air electrodes in both cases are exposed
to an aqueous electrolyte. The solid-state LiO2 battery may
function similarly to the aprotic system, although it is not
widely studied in detail yet due to the lack of a solid-state
electrolyte with sucient lithium ion conductivity. For this
reason, we only focus on the aprotic and aqueous LiO2
systems, with a particular emphasis on the former since it has
dominated the research eort on LiO2 batteries for the past
decade. It is assumed that the knowledge gained from these two
systems could also be applied to the solid-state and hybrid
systems due to the similarity of their chemistries.
Because oxygen is supplied as a reactant to the cell during
discharge, LiO2 cells dier from other batteries such as lead
acid, nickelmetal hydride, and lithium ion systems; they can
thus be constructed as a part of hybrid batteryfuel cell
systems. During electrochemical discharge, the Li electrode is
oxidized by releasing an electron to the external circuit to
produce Li ions in the electrolyte, whereas the oxygen is
reduced at a catalytic air electrode surface to form, in the case
of aqueous electrolytes, a lithium hydroxide product and, for
aprotic electrolytes, a lithium peroxide (Li2O2) or lithium oxide
(Li2O). This process is expected to be reversed on electrochemical charge in the aprotic system, making the cell
rechargeable. The LiO2 systems must be open with a porous
air electrode to allow the diusion of the oxygen gas to the
electrolyte/electrode interface. This design signicantly diers
from the conventional Li ion cell design with a completely
closed conguration. Use of a catalytic porous air electrode is
particularly important to aprotic LiO2 cells since it not only
facilitates the oxygen reduction and oxygen evolution reactions
but also provides the space to store the insoluble discharged
products, mainly lithium oxides. A porous air electrode with
highly catalytic activity toward oxygen evolution reaction is also
a key factor in realizing a secondary aqueous Liair battery.
This promising LiO2 battery technology is still in its infancy
and, no doubt, will require signicant research eorts in a
variety of elds to unlock its full potential. However,
researchers and industry alike see a great chance in its
development with the recent progress in advanced electrode,
electrolyte, and catalytic materials. Especially, the research eort
led by IBM to develop a Liair battery capable of driving a
commercial vehicle 500 miles on a single charge has sharply
accelerated research interest.35
Currently, many challenges prevent the realization of highperformance LiO2 batteries for the aprotic system. One of the
biggest is that the current aprotic electrolytes are often unstable
against several active discharge species (i.e., O2, O22 (Li2O2),
LiO2, and LiO2).55 Selection of the electrolyte is the key

Figure 1. Gravimetric energy densities (Wh/kg) for various types of


rechargeable batteries compared to gasoline. The theoretical density is
based strictly on thermodynamics and is shown as blue bars, while the
practical achievable density is indicated by orange bars and numerical
values. For Liair, the practical value is just an estimate. For gasoline,
the practical value includes the average tank-to-wheel eciency of cars.
Reprinted from ref 34. Copyright 2010 American Chemical Society.

density for LiO2 batteries is far less since it also strongly


depends on the porosity of the air electrode and the electrolyte
used at the air electrode side.50 In the case of an aprotic system,
the amount of insoluble discharge products stored in the air
electrode will ultimately determine the overall energy density,
while in the case of an aqueous system, the solubility of LiOH
in the base aqueous electrolyte is the limiting factor to the
energy density.34 On the basis of charge balance theory, Zheng
et al.50 predicted that the maximum possible gravimetric and
volumetric energy density of a LiO2 battery in an aqueous
electrolyte is considerably less than that when using an aprotic
electrolyte. Note that the usable energy density of gasoline for
automotive applications is approximately 1700 Wh/kg,
assuming an average tank-to-wheel eciency (12.6%) of the
U.S. eet. Fortunately, such energy density accounts only for
14.5% of the theoretical energy content of a fully charged Li
O2 battery, so it is not inconceivable that such a high energy
density may be achievable at the cell level, given intensive
research eort and long-term development.
Currently, four types of LiO2 batteries are under development and are designated by the type of electrolyte employed:
aprotic,10,51 aqueous,52 solid-state,53 and hybrid aqueous/
C

dx.doi.org/10.1021/cr400573b | Chem. Rev. XXXX, XXX, XXXXXX

Chemical Reviews

Review

Figure 2. Schematic cell congurations for the four types of Liair batteries.

Figure 3. Possible electrochemical and chemical reactions for LiO2 couples.

conducting membranes to protect the lithium electrode from


reacting vigorously with H2O. In the following sections, we
discuss the recent research to address the aforementioned
issues and share our perspectives with readers on how to better
understand the chemistries involved in the LiO2 systems and
improve their electrochemical performance.

component to improve the electrochemical performance of the


aprotic LiO2 system. Carbonate-based electrolytes have been
proved to be unstable in the aprotic LiO2 cells, while etherbased electrolyte showed some promise. The electrolyte
stability at the lithium electrode under the oxygen crossover
condition is also a big issue, necessitating development of new
electrolytes and a redesigned solidelectrolyte interface (SEI)
on the lithium electrode. Substantial diculties are faced in
preparing structures for the storage of insoluble discharge
products, presumably lithium oxide or lithium peroxide, at the
air electrode, which must also be made electrically conductive.
Porous carbon electrodes are the material of choice for most
current research, but pore clogging by the insoluble discharge
products in aprotic systems must be balanced with the need for
oxygen permeation. Catalysts have been shown to improve air
electrode performance, especially for lowering the charge
overpotential, with MnO2 being investigated most.56 However,
the mechanism of such improvement due to catalytic activity is
not yet clear, which could be the key factor in developing a
practical LiO2 battery. Signicant challenges have also been
faced at the pure lithium electrode: oxygen crossover eects57
and well-known dendritic lithium formation, long a problem in
lithium ion batteries, both of which eventually lead to decay of
the cell. Further complicating the design of LiO2 batteries is
the degradation of active materials by atmospheric contaminants, such as H2O and CO2,58 which require a selective
membrane that allows O2 permeation only. A dicult challenge
for aqueous LiO2 batteries is the development of good Li ion

3. APROTIC LIO2 BATTERY


3.1. Electrochemical Reactions in the Aprotic LiO2 System

As shown in Figure 2a, a typical aprotic LiO2 cell is composed


of a lithium electrode, an electrolyte consisting of dissolved
lithium salt in an aprotic solvent, and a porous O2-breathing
electrode that contains carbon particles and, in some cases, an
added electrocatalyst. Note that the oxygen reduction reaction
(ORR) during discharge and oxygen evolution reaction (OER)
during charge of a LiO2 cell occur at a three-phase boundary
involving the solid electrode, liquid electrolyte, and oxygen gas,
which makes the LiO2 system more complicated than the
conventional Li ion battery.
Assuming no parasitic side reactions are taking place, the
fundamental chemistry of a LiO2 electrochemical couple
during discharge will entail several possible reactions at the air
electrode via an oxygen reduction process, as described in
Figure 3. All the reactions except for reaction 2 (R2), the
disproportionation of lithium superoxide (LiO2), have a
standard redox potential close to 3.0 V vs Li/Li+, which
suggests that these reactions are equally favored thermodynamically. However, the oxygen reduction via a one-electron transfer
D

dx.doi.org/10.1021/cr400573b | Chem. Rev. XXXX, XXX, XXXXXX

Chemical Reviews

Review

(R1) would be most favorable kinetically during discharge,


which leads to the formation of LiO2. Subsequently, LiO2 can
turn into lithium peroxide (Li2O2) via either the disproportionation reaction (R2) due to its chemical instability or another
one-electron-transfer electrochemical process (R3). Upon
discharge, Li2O2 could also form on the air electrode through
a direct two-electron transfer, as described in reaction 4. It is
unclear at this moment which process is more favorable in
terms of the Li2O2 formation or whether they coexist and
compete with each other during the oxygen reduction, despite a
two-electron transfer having been reported recently.59 Although
the full reduction of O2 to lithium oxide (Li2O) through a
direct four-electron transfer (R6) is kinetically least favored for
the substrates with weak binding for oxygen (such as carbon)
considering that the oxygen dissociation on the substrates is
required, it is the most desired reaction for aprotic LiO2
batteries because of its higher specic energy and energy
density. However, if the cuto voltage is lowered to 2.0 V or
below (vs Li/Li+), Li2O forms as the discharge product on the
air electrode,6062 so discharge most likely takes multistep
reactions, for instance, R1 R3 R5, R1 R2 R5, or R4
R5. Due to the thermal stability and insulating nature of the
discharge product, the subsequent charge of Li2O cannot
proceed under the test condition.60 Therefore, Li2O2 is the
discharge product observed in most recent cell tests (cuto
voltage >2.0 V). In addition, from a kinetic point of view, the
formation of Li2O2 during discharge may be benecial to the
rate performance of the cell, since full cleavage of the OO
bond may not be necessary, given that a suitable catalyst can be
identied. In practice, the reactions that occur on the air
electrode during discharge are much more complicated due to
the strong interaction between the reduced oxygen species
(such as O2, O22 (Li2O2), LiO2, and LiO2) and other
components in the cell, namely, electrolyte, catalyst, and carbon
support, which directly aect the oxygen reduction reactions as
well as the subsequent oxygen evolution reactions. We further
review and discuss these issues in the following sections.
To make the LiO2 cell rechargeable if Li2O2 is the discharge
product, decomposition of Li2O2 to lithium and oxygen
(oxygen evolution reaction) is required, which could proceed
via either a two-electron process (Li2O2 2Li+ + 2e + O2) or
a one-electron process that involves the formation of LiO2
(Li2O2 Li+ + e + LiO2). Figure 4 shows a typical discharge
charge prole of an aprotic LiO2 cell, which shows a large
polarization and poor cycle life.44 One of the big challenges
here is to lower the charge overpotential to improve the
roundtrip eciency of the cell. Electrocatalysts, such as
manganese dioxide56 or noble metals,63 are often found to
have a positive eect on lowering the charge overpotential;
however, the exact role they play in the oxygen evolution
reaction is not clear. We discuss the parameters that have been
shown to have a signicant impact on the oxygen reduction and
evolution reactions and look ahead to how the electrochemical
performance of the LiO2 cell can be improved.

Figure 4. (a) Voltage prole of the rst cycle for a cathode containing
three-cycle atomic layer deposition (3c ALD) Pd in a carbon matrix
cycling in 1 M LiCF3SO3/TEGDME at 100 mA/g. (b) Cell capacity as
a function of the cycle number for the same air electrodes cycling in 1
M LiCF3SO3/TEGDME at 100 mA/g. The specic capacity and
current density were based on the active materials (carbon + Pd).
Reprinted from ref 44. Copyright 2010 American Chemical Society.

is not formed in aqueous systems. Cobalt pthalocyaninecatalyzed porous carbon was used as the air electrode to lower
the charge overpotential. Although researchers in this eld now
largely use liquid organic electrolytes in place of the polymer,
this work still remains seminal. Abraham and co-workers also
elucidated the mechanism of O2 reduction in LiO2 batteries
through the application of rotating disk electrodes.64
Read and co-workers6567 have largely focused on characterizing organic electrolytes for LiO2 batteries. They have shown
that electrolyte formulation has a very large eect on discharge
capacity and rate capability and that these performance
variables can be correlated to the solubility, diusion
coecients, and partial pressure of O2 in the electrolyte.66
The discharge product (Li2O2/Li2O) and location of the
product in relation to the air electrode are dependent on the
electrolyte type and discharge rate. It was subsequently shown
that the discharge product from oxygen reduction is Li2O if the
discharge potential is allowed to fall below 2.0 V (vs Li/Li+).
For this reason, many researchers limit the discharge window to
above 2.0 V, with the expectation that Li 2 O is too
thermodynamically stable to be easily decomposed upon
charging, as mentioned earlier. Read et al.68 have also studied
the eects of the air electrode porosity, claiming that, at low
cycling rates, the porosity of the air electrode dominates the

3.2. Early-Stage Research on the Aprotic LiO2 Battery

In 1996, Abraham and Jiang10 reported, for the rst time, an


aprotic LiO2 battery using a polymer organic electrolyte in
place of the aqueous electrolyte. In doing so, the battery was
able to be recharged, despite the large overpotential between
the discharge and charge as well as a very short cycle life. A
discharge product of Li2O2 was conrmed by Raman
spectroscopy and qualitative analysis; this discharge product
E

dx.doi.org/10.1021/cr400573b | Chem. Rev. XXXX, XXX, XXXXXX

Chemical Reviews

Review

Figure 5. Composite electrodes (Super P/-MnO2/Kynar) that contain the discharge products individually were subjected to charging in 1 M LiPF6
in PC under O2. (a) FTIR spectra of the as-prepared electrodes and the charged electrodes for each of the compounds, together with the spectrum of
a pristine electrode. (b) Corresponding charging curves at 70 mA/g. Since the theoretical capacities of the dierent compounds vary, to aid
comparison, the capacities are all normalized to unity (theoretical capacities: lithium propyl dicarbonate, 1000 mAh/g, 2e/mol; Li2CO3, 1500 mA
h/g, 2e/mol; CH3CO2Li, 750 mAh/g, 1e/mol; HCO2Li, 750 mAh/g, 1e/mol). (ce) MS gas analysis at the end of charging under O2 of
CH3CO2Li (c), C3H6(OCO2Li)2 (d), and HCO2Li (e). Note that unmarked peaks arise from fragments of CO2, H2O, O2, and Ar. (f) Gas evolution
measured by DEMS on oxidation of a composite electrode containing Li2CO3 in response to a stepwise increased current under Ar. The specic
capacity and current density were based on the mass of the carbon. Reprinted from ref 79. Copyright 2011 American Chemical Society.

and charge, electrocatalysts are required to achieve multiple


successful cycles in LiO2 batteries.
All the aforementioned work on the aprotic LiO2 batteries
is considered to be pioneering and, therefore, deserves
emphasis in this review. Since then, excitement over the
potential of LiO2 batteries has grown tremendously.
Researchers have gained much better understanding of the
LiO2 chemistry, although signicant challenges in the
development of advanced materials for the stable electrolytes
and electrodes have yet to be overcome.

battery performance. They developed a model that demonstrates how interparticle micro-, meso-, and macroporosities
dier in their importance to air electrode capabilities. Model
calculations suggest that much improvement can be achieved
with proper tailoring of the air electrodes physical properties.
The research in this eld was tremendously boosted in 2006
by Bruce and co-workers,51 who successfully tested a cell for
more than 30 cycles using an air electrode loaded with
electrolytic manganese dioxide as the electrocatalyst. This
performance was achieved despite the cell having a conventional carbonate-based electrolyte, which is now known to be
unstable to the oxygen reduction species (discussed in more
detail later). Bruce and co-workers also presented direct
evidence of oxygen release on cell charging, indicating the
reversibility of Li2O2 formation on discharge. In follow-up
work,56 Bruce et al. evaluated an array of dierent electrocatalysts and showed that the catalyst does not inuence the
discharge potential, but greatly aects the charging potential
and capacity retention per cycle. Overall, they unequivocally
showed that, despite the reactions occurring during discharge

3.3. Eect of the Electrolyte

One of the biggest challenges in the development of


rechargeable aprotic LiO2 cells is to develop a stable
electrolyte, in particular one that has an organic solvent and
can survive nucleophilic attack of the superoxide radical (O2),
an intermediate phase formed from the oxygen reduction
reaction upon discharge. Indeed, the choice of the electrolyte
might be the most critical factor in determining the
electrochemical performance of the cell. Although various
nonaqueous electrolytes have been investigated and used in Li
F

dx.doi.org/10.1021/cr400573b | Chem. Rev. XXXX, XXX, XXXXXX

Chemical Reviews

Review

Figure 6. (a) Reaction free energy prole for nucleophillic attack of O2 at the carbonyl and ethereal carbon atoms of PC and the reaction pathway
toward the formation of allyl alcohol (CH2CHCH2OH) and CO2 calculated within the continuum solvent model based on DFT. Reprinted from
ref 81. Copyright 2011 American Chemical Society. (b) Reaction energy prole for O2 attack on the ethereal carbon atoms of EC based on the
DFT and CCSD(T) levels of calculations. Reprinted from ref 80. Copyright 2011 American Chemical Society. (c) DFT results on the barriers for
activation of PC decomposition by other possible reduced O2 species, e.g., LiO2, LiO2, and Li2O2, besides O2. Reprinted from ref 55. Copyright
2011 American Chemical Society. (d) PC decomposition due to the reactivity of PC with a [100] Li2O2 surface through the nucleophilic addition to
the carboxylic carbon. Reprinted with permission from ref 82. Copyright 2011 Wiley-VCH.

ion batteries in the past few decades,69 these investigations are


normally performed under a closed and oxygen-free environment, which is less aggressive than in the LiO2 system. These
electrolytes should be re-examined for their stability in an
oxygen-rich environment, especially in the presence of the
oxygen reduced species, before they are tested in LiO2
systems. Besides stability against the superoxide radical, a
good electrolyte for LiO2 cells should also meet the following
criteria in addition to the typical requirements69 for electrolytes
used in Li ion batteries: (1) stability toward reaction with
lithium metal since this is the choice of the metal electrode for
LiO2 cells; (2) high boiling point and low volatility due to the
open cell system; (3) high oxygen solubility and diusivity to
facilitate the oxygen reduction and oxygen evolution reactions
on the air electrode; (4) low viscosity to improve the rate
performance of the oxygen electrode. Unfortunately, no single
electrolyte investigated so far meets these demanding requirements, despite extensive eorts having been devoted to that
end in the past few years. Understanding the reaction

mechanisms between the electrolytes and active oxygen


reduced species will, no doubt, be the key to developing a
stable electrolyte for LiO2 cells.
3.3.1. Lessons Learned from Organic CarbonateBased Electrolytes. Organic carbonate-based electrolytes, in
particular propylene carbonate (PC), have been widely
investigated in aprotic LiO2 batteries, mainly because they
are widely used in Li ion batteries. Propylene carbonate does
possess several advantages over other types of electrolytes, such
as a wide electrochemical window, low volatility, and a wide
liquid-temperature range.69 However, such electrolytes are not
stable in the presence of highly reactive species (i.e., O2, O22,
LiO2, and LiO2) from the oxygen reduction reaction, despite
the Li2O2 being claimed or assuming to be the discharge
product in previous research.56,70,71 In fact, Read et al.72
pointed out the possible instability of carbonate-based
electrolytes in LiO2 cells almost a decade ago. Even much
earlier than the rst report of the aprotic LiO2 battery in
1996,10 Aurbach et al.73,74 reported that the decomposition of
G

dx.doi.org/10.1021/cr400573b | Chem. Rev. XXXX, XXX, XXXXXX

Chemical Reviews

Review

carbonate was decomposed, consistent with previous ndings


by Bruce et al.79
In addition, the reactions of the electrolyte at lithium
peroxide surfaces could be responsible for the degradation of
the solvent. On the basis of a DFT study, Laino et al.82 revealed
that propylene carbonate could easily decompose on a [100]
Li2O2 surface with a barrierless reaction, as shown in Figure
6d. In contrast, the reaction by a Li2O2 molecular unit in
propylene carbonate solution has a barrier as high as 40 kcal/
mol among various reaction paths (e.g., hydrogen abstraction
and nucleophilic reactions).
At present, LiO2 cells with carbonate-based electrolytes
cannot be considered viable, as the overall reaction is the
continuous and irreversible oxidation of the electrolytes, and
therefore, the battery is in principle irreversible. The lesson
learned from this case is that direct implantation of a wellestablished electrolyte for the Li ion battery into the LiO2
system was not successful. On the other hand, the lessons
learned from the study of carbonate-based electrolyte have
provided a profound understanding of the operation of aprotic
LiO2 batteries and resulted in a more reliable process for
seeking stable electrolytes. It should be pointed out that the air
electrode catalysts investigated with organic carbonate-based
electrolyte need to be re-examined, since the claimed catalytic
activities toward the oxygen reduction and oxygen evolution
reactions appear to be more related to the decomposition of the
electrolyte than the reversible formation and decomposition of
Li2O2.
3.3.2. Ether-Based Electrolyte. After realizing that the
carbonate-based electrolytes have suered from severe
decomposition during the operation of LiO2 cells, researchers
have focused much attention on the experimental55,59,60,78,8391
and theoretical80,81,9295 search for other solvents that are
stable against attack by reduced O2 species and can avoid other
possible decomposition problems. Ether-based solvents, such as
1,2-dimethoxyethane (DME) and tetraethylene glycol dimethyl
ether (tetraglyme or TEGDME), have attracted signicant
attention for the LiO2 battery very recently, mainly due to the
relatively high stability with respect to superoxide radicals and
oxidation potentials (>4.5 V versus Li/Li+), inammability,
high thermal stability, and low cost. In fact, Read67 rst
reported that ether-based electrolytes perform well in the Li/O2
cell, showing both good stability and excellent rate capability,
even before the recognition of the instability issue of carbonate
solvents. Read found that the viscosity of ether-based
electrolytes is lower than that of carbonate-based electrolytes,
while still having similar oxygen solubility. Read also pointed
out that the viscosity becomes the key factor in determining the
discharge capacity and rate capability of the cell once a certain
level of oxygen solubility is reached. Unfortunately, Read did
not show any evidence for the formation of Li2O2 upon
discharge of the cell, although we know now that the discharge
product using ether-based electrolytes is indeed Li2O2.
Recently, several groups have found evidence that the major
discharge product is Li2O2 when ether-based solvents are used
in the LiO2 cell. McClosky et al.88 investigated the stability of
DME in a LiO2 cell containing dierent air electrode catalysts,
where they conrmed that Li2O2 is the dominant product of
the electrochemistry. Interestingly, they found no catalytic
activity compared to pure carbon when Au, Pt, or MnO2
nanoparticles were used as the air electrode catalyst. In their
follow-up paper, McClosky et al.96 showed that the amount of
O2 evolved upon charging accounts for only 60% of the

propylene carbonate could occur on the intermediates/


products of the electrochemical reduction of oxygen.
Unfortunately, researchers did not pay much attention to the
instability of carbonate-based electrolytes until 2010, when
Mizuno et al.75 presented direct evidence that the discharge
products of the PC-based LiO2 cell were mainly lithium
carbonate (Li2CO3) instead of the desired Li2O2. Since then, a
wealth of experimental and theoretical evidence has indicated
that the organic carbonates (e.g., propylene carbonate, ethylene
carbonate, and dimethyl carbonate) are not stable toward the
oxygen reduction products.55,7581 For example, Bruce and coworkers79 demonstrated that the discharge products from the
LiO2 cell containing alkyl carbonate electrolyte mainly consist
of various carbonate species, including lithium formate, lithium
acetate, and lithium carbonate, with no evidence for the
formation of Li2O2 from Fourier transform infrared (FTIR)
spectra, as shown in Figure 5. These carbonate compounds can
be oxidized on charging (Figure 5),79 leading to the formation
of CO2 as the primary charge product, as evidenced by in situ
dierential electrochemical mass spectrometry (DEMS)
combined with 18O isotope labeling experiments.78 This charge
product is responsible for the large overpotential, as widely
observed by other researchers. As a net eect, the cycle
performance of the LiO2 cell using carbonate-based electrolytes is associated with the consumption of the electrolytes
rather than reversible discharge/charge of the desired Li2O2.
The intermediates from the oxygen reduction reaction during
the discharge of the cell are believed to be responsible for the
decomposition of the carbonate-based electrolytes. To understand the nature of the reaction mechanism, a density
functional theory (DFT) study and coupled-cluster calculations
by Bryantsev and Blanco81 found that propylene carbonate is
highly susceptible to the nucleophilic attack of superoxide
radical (O2), as shown in Figure 6a. Subsequent reactions led
to the formation of carbonate species as the discharge products.
Bryantsev and Blanco81 also found that similar decomposition
pathways are valid for other organic carbonate solvents, e.g.,
ethylene carbonate (EC) and dimethyl carbonates (DMC),
which have computed activation barriers of about 12.415.5
kcal/mol. According to Bryantsev and Blanco,81 the nucleophilic attack by O2 at ethereal carbon atoms is a common
mechanism of degradation of organic carbonate solvents
(Figure 6). However, other oxygen reduced species, such as
O2, O22, LiO2, and LiO2, may exist in the LiO2 cell, and
they could also cause the electrolyte decomposition.
From the modeling results based on DFT and high-level
quantum chemistry calculation, Zhang et al.55 demonstrated
that the decomposition of propylene carbonate starts from
ring-opening, i.e., CO bond breaking, which is consistent
with the ndings of Bryantsev et al.81 The calculated energy
barriers of the CO bond breaking for all four oxygen reduced
species considered (i.e., O2, O22, LiO2, and LiO2) are quite
small, with LiO2 surprisingly being the most reactive, as shown
in Figure 6c. Once the ring of PC is opened, the subsequent
CO bond breaking is easier and thermodynamically favorable;
this condition could lead to the formation of Li2CO3 or other
products such as formaldehyde and acetaldehyde with the
assumption that a second electron transfer can occur. This
theoretical prediction was further conrmed by X-ray photoelectron spectroscopy (XPS),55 which indicated that lithium
carbonate was the major product formed during discharge, with
only minor amounts of lithium oxides being formed. The XPS
data also indicate that, during the rst charge cycle, the lithium
H

dx.doi.org/10.1021/cr400573b | Chem. Rev. XXXX, XXX, XXXXXX

Chemical Reviews

Review

consumption of O2 during discharge in the same electrolyte,


according to the in situ quantitative gas-phase mass
spectrometry (DEMS) measurements shown in Figure 7.

Figure 8. (a) Optimized structure for the (Li2O2)16DME cluster


(quintet) and (b) transition-state structure for the abstraction of a
primary hydrogen from the DME at the B3LYP/6-31G(d) level of
theory. Selected bond distances () and relative energies (eV) are also
shown. Reprinted from ref 93. Copyright 2013 American Chemical
Society.

unpaired spin on a LiOLi site (if present on the surface)


may be favorable for decomposition of DME, which likely leads
to the formation of reactive radicals and, subsequently, in the
presence of oxygen, can lead to oxidized species such as
aldehydes and carboxylates as well as LiOH on the surface of
the lithium peroxide. Other groups also employed DME as the
baseline electrolyte for the evaluation of the air electrode
catalysts.27,87,97107 Although DME is more resistant to the
superoxide radicals, the practical application of DME in LiO2
cells with long-term operation will be restricted due to its high
volatility.58
Longer chain ether solvents such as TEGDME have been
considered as another choice for the electrolytes in aprotic Li
O2 batteries.46,47,57,60,91,95,108135 In general, longer chain ethers
with larger molecular weights are stable in contact with Li metal
and have less volatility and polarity. Moreover, TEGDME was
shown to be stable against superoxide species by Black et al.,135
who chemically generated metastable solvated LiO2 from the
reaction of KO2 with dicyclohexyl-18-crown-6 (crown ether)
and then LiPF6 in solution, as shown in Figure 9. The
metastable solvated LiO2 may also exist during the cell
operation through the combination of an electrochemically
generated superoxide radical (O2) from the oxygen reduction
reaction with the solvated Li+ in the electrolyte, although it
likely has a nite lifetime in solution.60 Interestingly, Black et
al.135 observed the decomposition of poly(vinylidene diuoride) (PVDF), one of the widely used binders in LiO2 and Li

Figure 7. Gas evolution from cells employing DME: (a) discharge


charge voltage curves and corresponding (b) O2 and (c) CO2
evolution during charging of cells using various air electrode catalysts.
Reprinted from ref 96. Copyright 2012 American Chemical Society.

They attributed this behavior to the existence of some weak


reactions between Li2O2 and DME under the cell potential.
The in situ DEMS technique coupled with galvanostatic cycling
provides a very powerful tool to analyze gas evolution and
determine the reversibility of Li2O2 formation/decomposition
during the operation of LiO2 cells. Recently, Assary et al.93
showed that, on the basis of DFT calculation results, the
interactions of DME with Li2O2 clusters have a high possibility
of such reactions, which could be responsible for the
degradation of the solvent in addition to the O2 that has
been postulated to be present in the electrolyte. Small clusters
of up to 16 Li2O2 units were modeled to investigate the
reactions between ether (DME) and various possible surface
sites on the Li2O2 discharge products,93 as shown in Figure 8.
The computations indicated that hydrogen abstraction by an
I

dx.doi.org/10.1021/cr400573b | Chem. Rev. XXXX, XXX, XXXXXX

Chemical Reviews

Review

Figure 9. Exploratory reactions of superoxide that parallel those in a


LiO2 cell: (A) KO2 + LiPF6/TEGDME; (B) KO2 + LiPF6/
TEGDME + carbon; (C) electrochemical cell (carbon catalyst); (D)
KO2 + LiPF6/TEGDME/PVDF; (E) KO2 + LiPF6/TEGDME/
PVDF/-MnO2 catalyst. KO2 = KO2 (crown ether), and PVDF =
poly(vinylidene diuoride). Reprinted from ref 135. Copyright 2013
American Chemical Society.

ion batteries, through dehydrouorination along with the


formation of H2O2 by the chemically generated LiO2. Black et
al.135 also showed that, in the absence of binder and catalyst,
the crystallization of Li2O2 is governed by its rate of nucleation/
precipitation from LiO2, solubility, and interaction with the
carbon surface. This work provided vital insight into the
reactivity of superoxide radical with materials used in the LiO2
system, which should be seriously considered as a screening
method in searching for stable electrolytes for LiO2 batteries.
Jung et al.91,136 reported an improved LiO2 battery using
TEGDME/LiCF3SO3 electrolyte, which exhibited high capacity
and stable cycling performance with an impressive rate
capability. The reversible formation of Li2O2 was conrmed
by time-of-ight secondary ion mass spectrometry (TOFSIMS), along with scanning and transmission electron
microscopy, indicating morphological reversibility of semicrystalline Li2O2 particles. Moreover, Li2CO3 was not detected
in the discharge and charge states, showing that TEGDME/
LiCF3SO3 electrolyte is stable during cell operation. This Li
O2 cell showed very stable cycling performance up to 100
cycles with a capacity of 1000 mAh/gcarbon at both a low rate
(50 mA/gcarbon) and a high rate (1000 mA/gcarbon). Even under
a very high capacity of 5000 mAh/gcarbon, this LiO2 cell is
capable of operating over 30 cycles at a rate of 500 mA/gcarbon,
as shown in Figure 10.
The reactivity of a new silicon-containing oligo(ethylene
oxide) solvent, (triethylene glycol-substituted methyl)trimethylsilane (1NM3), toward the oxygen reduced species
has also been investigated because of its good properties as an
electrolyte, such as low glass transition temperature, eective
ionic transport, low viscosity, and good lithium ion
conductivity.57,137140 On the basis of the DFT calculation,
Zhang et al.55 found that this solvent seems to be more stable
to the highly active oxygen reduction species when compared
to propylene carbonate. The XPS data collected after a single
discharge show that only lithium oxides are formed when
1NM3 is used in place of propylene carbonate in a LiO2 cell.
Moreover, the formed Li2O2 decomposed upon charging at a
low charge potential (3.5 V vs Li/Li+), as shown in Figure 11. It
seems that the increased stability of the 1NM3 solvent results

Figure 10. Cycling performance of the lithiumTEGDME/


LiCF3SO3O2 battery under various specic capacity limits and
current densities: (a) voltage proles of 100 cycles with a capacity limit
of 1000 mAh/g at a current density of 50 mA/g; (b) voltage proles
of 100 cycles with a capacity limit of 1000 mAh/g at a current density
of 1000 mA/g; (c) voltage proles of 30 cycles with a capacity limit of
5000 mAh/g at a current density of 500 mA/g. The specic capacity
and current density were based on the mass of the carbon. Reprinted
with permission from ref 91. Copyright 2012 Nature Publishing
Group.

Figure 11. First charge and discharge cycles of a LiO2 cell with PC
and 1NM3 electrolyte. The current density applied in both cases is 100
mA/g, based on the mass of the carbon + catalyst. Reprinted from ref
55. Copyright 2011 American Chemical Society.
J

dx.doi.org/10.1021/cr400573b | Chem. Rev. XXXX, XXX, XXXXXX

Chemical Reviews

Review

in a signicantly lower overpotential for the LiO2 cell, which


could be due to the oxidation of lithium oxides being easier
than the oxidation of carbonate formed in the propylene
carbonate cells. However, the cycle performance of the cell
using 1NM3 electrolyte is still poor, indicating that the longterm stability of 1NM3 solvent is problematic. The electrolyte
solvent stability plays a key role in the performance of LiO2
batteries and will be a key factor in improving their eciencies.
Bruce and co-workers83 investigated the stability of linear or
cyclic ether-based electrolytes (e.g., tetraglyme, triglyme,
diglyme, 1,3-dioxolane, and 2-methyl-THF) in LiO2 cells by
combining electrochemical measurements with powder X-ray
diraction (XRD), FTIR spectroscopy, and nuclear magnetic
resonance (NMR) spectroscopy. In all cases, Li2O2 was the
oxygen reduction product on the rst discharge, conrming
that ether-based electrolytes are more stable than carbonatebased ones against oxygen reduced species. The ether solvents
also decomposed during the operation of the cell accompanied
by the formation of Li2O2. For instance, Bruce et al.83 found
that the decomposition products for linear-chain ethers consist
of a mixture of Li2CO3, HCO2Li, CH3CO2Li, polyethers/esters,
CO2, and H2O. Their cycling studies on TEGDME indicated
that the proportion of Li2O2 diminishes on cycling in favor of
greater electrolyte decomposition, as evidenced by the XRD
and FTIR measurements shown in Figure 12. On the basis of
these results, they concluded that ether-based electrolytes are
not suitable for LiO2 cells. We point out, however, that the
eect of the salt was not taken into account in their study. In
fact, LiPF6, which was used in their ether-based electrolytes as
the lithium salt, is suspected to be stable to Li2O2. Moreover,
LiPF6 has been known to decompose to form HF in the
presence of H2O, even with a trace amount. As a consequence,
the observed decomposition of ether solvents may result from
the active species generated from the side reaction of the
lithium salt. When LiCF3SO3 replaces the LiPF6 in the
TEGDME electrolyte, the cell performance is signicantly
improved, as demonstrated by Jung et al.91,136 The eect of
lithium salt on the cell performance is further discussed later.
3.3.3. Other Organic Electrolytes. Several other types of
organic solvents have also been investigated in aprotic LiO2
systems, including acetonitrile (ACN),84,96,141 dimethyl sulfoxide (DMSO),59,85,99,142148 dimethylformamide (DMF),149
triethyl phosphate,90 N,N-dimethylacetamide (DMA),150,151 Nmethylpyrrolidone (NMP),34 methoxybenzene,152 and ionic
liquids,86,153,154 some of which demonstrated relatively high
stability toward oxygen reduction intermediates (LiO2) and
products (Li2O2) in a LiO2 cell. One of the particularly
interesting solvents is DMSO. Peng et al.59 reported a very
steady rechargeable LiO2 cell using LiClO4/DMSO as the
electrolyte coupled with a nanoporous gold (NPG) air
electrode, which is able to cycle up to 100 times with little
decay of the capacity, as shown in Figure 13. The FTIR, Raman
spectroscopy, NMR, and DEMS results indicated that Li2O2 is
formed at the air electrode with >99% purity, even on the 100th
cycle, and it is completely oxidized on charge. The charge-tomass ratio on discharge/charge was determined to be 2e/O2
by the DEMS results, suggesting that Li2O2 formation/
decomposition is the overwhelming reaction during the
operation of the cell. Note that the charge potential is still
high (up to 4.0 V), although a low charge plateau occurs at 3.3
V. Charge transfer between the Li2O2 particles and the solid
electrode surface is likely responsible for the voltage polarization on charge because of the insulating nature of Li2O2,

Figure 12. (a) XRD patterns and (b) FTIR spectra of O2 electrodes
cycled in 1 M LiPF6/TEGDME electrolyte. Reprinted with permission
from ref 83. Copyright 2011 Wiley-VCH.

Figure 13. Charge/discharge curves (left) and cycling prole (right)


for a LiO2 cell with a 0.1 M LiClO4/DMSO electrolyte and an NPG
air electrode at a current density of 500 mA/g (based on the mass of
Au). Reprinted with permission from ref 59. Copyright 2012 Science.

which is a signicant problem facing the aprotic LiO2 battery.


Very recently, the same group147 seems to have solved the issue
of large charge overpotential in their previous cell by
introducing a redox mediator, tetrathiafulvalene (TTF), to
the DMSO electrolyte. During cell charging, TTF is oxidized to
TTF+ at the air electrode surface, which subsequently oxidizes
K

dx.doi.org/10.1021/cr400573b | Chem. Rev. XXXX, XXX, XXXXXX

Chemical Reviews

Review

Figure 14. Cycling stability of LiO2 air electrodes that employ a redox mediator. Load curves (constant-current discharge/charge) for the 1st, 20th,
and 100th cycles of a cell with 1 M LiClO4 in DMSO that contained 10 mM TTF at a nanoporous gold electrode under O2, with rates increased
from 0.078 mA/cm2 (a) to 0.196 mA/cm2 (b), to 0.313 mA/cm2 (c), and to 1 mA/cm2 (d). The similarity of the load curves with cycle number
demonstrates the stability of the Li2O2 formation/decomposition on cycling in the presence of TTF-mediated oxidation of Li2O2 on charging.
Reprinted with permission from ref 147. Copyright 2013 Nature Publishing Group.

Abraham and Jiang.10 Using a poly(ethylene oxide)/lithium


triate (PEO/LiCF3SO3) solvent-free polymer electrolyte,
Hassoun et al.165 were able to demonstrate a highly reversible
LiO2 reaction with low redox overpotential and the close to
100% charge ratio between the oxygen reduction and oxygen
evolution reactions via the potentiodynamic cycling with
galvanostatic acceleration (PCGA) and cyclic voltammetry
(CV) analyses on the solid-state cell. Such solid-state
electrolyte is a very good medium for the study of the
electrochemistry of the LiO2 couple in water-free lithium
oxygen cells. They showed that, at low current densities, the
perfect deconvolution of the peaks in the PCGA allows
separation of at least two electrochemical processes attributed
to peroxide and normal oxide. However, the challenge facing
these solid-state electrolytes is to improve their ionic
conductivity and manage the huge volume variation of the air
electrode between discharge and charge. With a recent
breakthrough in the solid-state electrolytes,166 which shows
even higher ion conductivity than that of liquid aprotic ones,
the utilization of such solid electrolytes in the aprotic LiO2
system could possibly be more feasible.
Mixed-solvent electrolyte probably represents a new research
direction for identifying stable electrolytes for aprotic LiO2
batteries, although they have not been widely investigated
yet.167 In Li ion batteries, however, the mixed-solvent
electrolytes are very popular and perform much better than
the individual components. A good example is the mixed
carbonate electrolyte EC/DMC,168,169 in which the electrochemical stability can reach up to 5.0 V vs Li/Li+; without EC,
pure DMC can be oxidized only to about 4.0 V vs Li/Li+. The
synergetic eect from the mixed solvent appears to improve the
electrochemical performance of the electrolytes in the Li ion
batteries. By the same token, the performance of the LiO2 cell
could also be enhanced by blending two or more organic

the solid Li2O2 while TTF+ turns back to TTF. Therefore, the
redox mediator acts as a molecular electronhole transfer agent
on charging, which permits ecient oxidation of solid Li2O2. As
a consequence, the cells that contain TTF were able to operate
for 100 cycles with complete reversibility of Li2O2 formation/
decomposition on each cycle, even under a very high current
density (1 mA/ggold), as shown in Figure 14. Redox mediators
(often known as redox shuttles) have also been widely
demonstrated for overcharge protection in lithium ion
batteries.155161 Note that DMSO is not stable with bare
lithium; therefore, a protected Li electrode is required for
operation of the LiO2 cell in DMSO electrolytes. In this case,
a perfectly protected Li electrode could be obtained by simply
dipping lithium into the LiClO4/PC electrolyte, according to
Bruce et al.59
The O 2 electrochemistry of aprotic LiO 2 cells in
acetonitrile has also been investigated by several groups. Laoire
et al.141 reported that, in acetonitrile-based electrolyte, Li2O2
formed as the oxygen reduction product on the air electrode in
a stepwise fashion (O2 O2 O22), which was conrmed
by Peng et al.84 In addition, Laoire et al.141 found that O2 is
much more dicult to reduce in acetonitrile than in DMSO,
and McCloskey et al.96 revealed that a LiO2 cell with carbon
as the air electrode exhibited a very high reversibility of Li2O2
formation/decomposition when acetonitrile was employed as
an electrolyte. All of these results demonstrated that acetonitrile
is relatively stable toward the oxygen reduction species.
However, the practical application of acetonitrile in aprotic
LiO2 systems is restricted because of its high volatility, similar
to that of DME.
Solid-state electrolytes have also been reported and have
shown some promise in aprotic LiO2 batteries.53,162164
Noteworthy, the rst rechargeable aprotic LiO2 cell was built
up on the basis of a solid-state electrolyte, namely,
polyacrylonitrile-based polymer electrolyte, as reported by
L

dx.doi.org/10.1021/cr400573b | Chem. Rev. XXXX, XXX, XXXXXX

Chemical Reviews

Review

solvents, given that they are stable (at the air electrode and
lithium electrode) during the operation of the cell.
3.3.4. Eect of Lithium Salt. Besides the stability of
solvents, another source of detrimental electrolyte decomposition in LiO2 batteries could be from side reactions due to
the decomposition of lithium salt, an indispensable component
of the electrolyte. In contrast to the extensive eorts on solving
the instability issue of solvents, the eect of lithium salts on the
stability of the electrolytes in the aprotic LiO2 system has not
been investigated in detail yet, even though the choice of
lithium salts is quite limited. The lithium salts that have been
employed in the aprotic LiO2 system include LiPF6, lithium
bis((triuoromethyl)sulfonyl)imide (LiTFSI), LiClO 4 ,
LiCF 3SO3, LiBF4, LiCl, and lithium bis(oxalato)borate
(LiBOB), with the rst four being used the most. Analogous
to the Li ion battery, a good lithium salt for the LiO2 battery
has to meet certain requirements, namely, (1) the salt should
have high enough solubility in the solvent to support the ion
transport and (2) the anion of the salt should be inert to the
solvent and other cell components, such as current collectors
and separators. In addition to these basic requirements, the salt
used in the LiO2 system should be inert to the aggressive
oxygen reduction species, such as Li2O2 and O2 radical.
Recent evidence has shown that some lithium salts are not
stable toward Li2O2 formed in the LiO2 cells, with LiPF6
being the most obvious one. The XPS results of Oswald et
al.170 indicated an almost immediate formation of LiF if the
Li2O2 is exposed to a LiPF6-containing electrolyte. The
formation of LiF is partially due to the presence of HF,
which is generated by uorine from uorinated anion (PF6)
and protons from water impurities in the electrolyte. The XPS
depth proles also revealed another possible reason for the
formation of LiF, the direct reaction of Li2O2 and LiPF6. By
analyzing the discharge products of aprotic LiO2 cells in
TEGDME-based electrolytes with various salts (LiPF6, LiClO4,
LiTFSI, and LiCF3SO3) using Raman, XPS, and NMR
techniques, Veith et al.171 provided strong evidence that the
discharge product in ether-based electrolytes contains a certain
level of insoluble Li salt, although the major discharge product
is Li2O2. They believe that the discharge chemistries of LiO2
cells appear to undergo similar decomposition of the lithium
salt, resulting from the strong interaction of the salt and the
reactive O2 or reduced oxygen species, such as O2.
As discussed above, the stability of lithium salt, especially in
the presence of reduced oxygen species (O2) and a trace
amount of H2O, may signicantly aect the cyclability and
capacity of LiO2 cells. Applying a combined experimental and
computational approach, Du et al. provided evidence that the
stability of the electrolyte in the LiO2 cell strongly depends
on the compatibility of the lithium salt with the solvent.172 In
the case of the LiPF6/1NM3 electrolyte, they found that HF
formed from decomposition of LiPF6 triggers the decomposition of the 1NM3 solvent, as schematized in Figure 15.
These reactions lead to severe degradation of the electrolyte
and cause the poor cyclability of the cell. The same reactions
are not observed when LiTFSI or LiCF3SO3 is used as the
lithium salt in 1NM3 solvent, suggesting that the stability of the
electrolyte in Liair cells depends on the compatibility of the
lithium salt with the solvent.
Although not yet having received much attention, the
possible decomposition of the lithium salt and the subsequent
eect on the stability of the electrolytes and performance of the
aprotic LiO2 cells should not be neglected. The challenge in

Figure 15. Schematic reaction mechanism for decomposition of


LiPF6/1NM3 electrolyte in the LiO2 cell. Reprinted with permission
from ref 172. Copyright 2013 PCCP Owner Societies.

investigating the stability of lithium salts in LiO2 cells,


however, strongly relies on a stable solvent, which, in turn,
should have the capability to dissolve a certain amount of
various lithium salts. This situation suggests that progress in
developing a stable electrolyte for the aprotic LiO2 system can
only be made in a stepwise fashion.
3.3.5. HardSoft AcidBase (HSAB) Theory. To design
a rational approach to the selection of organic electrolyte for
the rechargeable aprotic LiO2 battery, a more fundamental
theory or model is required to elucidate the inuence of
solvent/salt on O2 electrochemistry. Laorie et al.141 successfully
adopted the concept of HSAB theory to explore various
electrolyte systems to determine the eect of a solvents donor
number (DN) and acceptor number (AN) on the discharge
products of LiO2 cells. In Pearsons HSAB theory,173 Lewis
acids and bases are further classied into hard and soft
subcategories, where hard acids tend to interact with hard bases
while soft acids tend to interact with soft bases. In general, hard
acids/bases have relatively small ionic radii and are dicult to
polarize and vice versa.
According to the HSAB theory, Li+ is a hard Lewis acid and
has a high anity for a hard Lewis base, such as peroxide
(O22) or monoxide (O2). However, in the electrolyte
solution, the hard Lewis acid Li+ ions are normally solvated
by the solvents to form a complex (Li+(solvent)n), which will
then modulate the acidity of Li+ dependent on the DN of the
solvents. Laorie et al.141 showed that the bond strength of Li+
solvent in the complex would follow the solvent DN scale as
DMSO > MeCN > DME > TEGDME. On the other hand, the
superoxide (O2), one of the possible oxygen reduction species
in the LiO2 system, tends to be a moderately soft Lewis base
because of its relatively large radius and low charge density.
Therefore, the superoxide anion has low anity for the hard
Lewis acid Li+ in the solutions, for instance, TEGDME/LiPF6
electrolyte. Consequently, the superoxide anion formed as the
rst reduction product, if it does form, will prefer either to
decompose via a disproportionation reaction (assuming the
solvent is stable to superoxide anion) or to undergo a fast
second reduction, both of which will lead to the formation of a
hard Lewis base, peroxide anion (O22). Once the peroxide
anion forms, it will then combine with the hard Lewis acid Li+
to produce a stable O2 reduction product, Li2O2, on the basis of
the HSAB theory. This could explain the observation by many
other researchers of Li2O2 formation as the dominant discharge
M

dx.doi.org/10.1021/cr400573b | Chem. Rev. XXXX, XXX, XXXXXX

Chemical Reviews

Review

superoxide anion against ethers is partially dependent on the


potentials applied during cell discharge and charge, which is
certainly related to the electrocatalysts on air electrodes.

product in the LiO2 cells using TEGDME-based electrolytes.47,91,120,124,127,174,175 Theoretically, Li2O, the ultimate
reduction product of O2, should also be able to form as a
stable discharge product in a LiO2 cell, considering that the
monoxide (O2) anion is an even harder Lewis base. However,
the charge of Li2O cannot proceed under the test condition due
to the need for 4-electron transfer, which is kinetically
unfavorable in the electrochemical process; therefore, the
discharge product is normally limited to Li2O2 by controlling
the cuto discharge voltage of the LiO2 cell to above 2.0 V vs
Li/Li+.
As mentioned above, the formation of Li+(solvent)n
complexes would modulate (lower) the acidity of Li+,
determined by the DN of the solvent. Therefore, the
superoxide anions could be stabilized to a certain degree
before they transform to O22 via a chemical or electrochemical
pathway, because of the increased anity between O2 (soft
Lewis base) and the solvated Li+ with lower acidity (toward soft
Lewis acid). This could explain the distinct O2/O2 couple seen
in the DMSO/LiPF6 electrolyte.59,85,143147 Thus, the HSAB
theory could provide guidance to further modulate the Li+
acidity in the electrolytes to form dierent types of discharge
products during the oxygen reduction reaction. Successful
implementation of this theory will be benecial for future
investigation of dierent electrolytes for optimization of the
LiO2 cell.
3.3.6. Brief Summary. In summary, the organic electrolytes
(both solvents and lithium salts) play the most critical role in
an aprotic LiO2 cell and determine whether a truly
rechargeable LiO2 battery can be realized. Numerous studies
showed that the stability of the electrolytes during the oxygen
reduction and oxygen evolution processes is the key challenge
for the aprotic LiO2 cell. With no doubt, searching for fully
stable electrolytes in the oxygen-rich electrochemical environment is the research priority at present. Carbonate-based
electrolytes have proved to be highly unstable toward the
oxygen reduction species. However, there is still a large amount
of research work using carbonate-based electrolytes to
investigate the catalytic activities of the air electrode
materials,126,133,176216 despite the fact that the severe
instability of these electrolytes has been reported. The catalytic
activity of these air electrode materials needs to be re-examined
in more stable electrolytes. Ether-based electrolytes appear to
be relatively stable in the presence of the reduced oxygen
species, as evident by the reversible formation/decomposition
of Li2O2; however, their electrochemical behavior, in particular
on charge and during long-term cycling, remains to be
thoroughly investigated for optimization of the LiO2 cell.
Both DMSO59,147,217 and TEGDME91,136 are probably the
most promising organic solvents under investigation. The best
electrochemical performance (including cycle performance) has
been achieved with LiO2 cells having these electrolyte
solvents. Lithium salt, an indispensable component of the
electrolyte, deserves much more attention since there is
evidence it has an eect on the stability of the electrolytes in
LiO2 cells. Design of a robust strategy for eectively screening
the stability of various electrolytes would be greatly benecial to
the development of a LiO2 battery for practical application. In
addition, we noticed that there are several dierent or even
conicting reports on the reactivity of superoxide anion (O2)
against ethers, which certainly requires deep fundamental
understanding using the state-of-the-art characterization
techniques. In our opinion, however, the reactivity of the

3.4. Electronic and Magnetic Properties of Li2O2 and


Relevance to the LiO2 Battery

Another research area that deserves attention is the eect of the


electronic structure and magnetic properties of Li2O2 on the
electrochemical performance (particularly the charge overpotential) and electrolyte stability of LiO2 cells. Better
understanding of these properties could help us elucidate the
charge and discharge chemistries involved in the LiO2 battery.
In general, the most thermodynamically stable structures in
the condensed phase for known stoichiometric Li x O y
compounds are Li2O (lithia) and Li2O2,218 both of which are
theoretically predicted to be semiconductors with a high
electronic band gap; for instance, bulk Li2O2 shows a band gap
of 4.9 eV.219221 Since high electronic conductivity and
charge transfer are essential for any redox chemistry in the
battery system, such a high band gap fails to explain the
reversible formation of Li2O2 on the air electrode during
discharge/charge of a LiO2 cell, especially under very high
current density.91,136 Recently, Hummelshoj et al.219,222
proposed a theory based on metallicity that shows the
possibility of lowering the electronic band gap by inducing
some Li vacancies into bulk Li2O2. Depending on the vacancy
concentration, the formation energy of Li2O2 could be reduced
to 3.0 eV, which will enable reasonable electron and charge
transfer in Li2O2. The theory is supported by the nding that
Li2O2 observed as the discharge product of a LiO2 cell is
generally in the form of nanoparticles,61,221 which are known to
have vacancies and defects on the surface.
Since neither the mechanism of formation and growth of
Li2O2 during discharge nor the mechanism of the decomposition of Li 2 O 2 during charge is well understood
experimentally, computational studies on this topic will likely
provide valuable insight in understanding the correlation
between the reversible formation of Li2O2 and its electronic
and magnetic properties, if any. At the current stage, theoretical
calculation on small Li2O2 clusters is the choice of these studies,
considering that computations on large clusters or particles
would be much more complicated and costly. The information
gathered from theoretical computations on small Li2O2 clusters
could provide insight into the properties of larger nanoparticles
as well as their nucleation during cell discharge.
In a recent paper, Lau et al.223 reported on a DFT study of
the structure of (Li2O2)n clusters, where many possible
structures of the clusters and the corresponding electronic
states were considered. They found that, surprisingly, the triplet
state is signicantly stabilized relative to the singlet in these
clusters, especially for clusters larger than the dimer. Their DFT
calculations also showed the existence of a high spin state in a
large Li2O2 cluster with n = 16, which can be characterized by
OO pairs protruding from the surface but still chemically
bonded to the Li of the clusters, as shown in Figure 16. These
Li2O2 clusters have shorter OO distances and a localized
unpaired spin on the surface compared to a peroxide pair found
in a Li2O2 bulk crystal, which probably suggests the existence of
superoxide-like surface structures. This theoretical nding could
explain the ferromagnetism with magnetic moments of Li2O2,
earlier predicted by Radin et al.224 using rst-principles
calculations. More importantly, if these superoxide-like surface
structures indeed exist experimentally, they could have
N

dx.doi.org/10.1021/cr400573b | Chem. Rev. XXXX, XXX, XXXXXX

Chemical Reviews

Review

peroxide with superoxide-like properties characterized by a lowtemperature magnetic phase transition (49.7 K) and a high O
O stretching frequency (Raman peak at 1125 cm1). A low
charge potential (3.3 V) was also observed in a LiO2 cell using
1NM3/LiTFSI as the electrolyte, where the Li2O2 discharge
product shows magnetic behavior.225 These experimental
ndings suggest that a superoxide surface species lowers the
charge potential, and therefore, control of the forms of Li2O2
produced on discharge is important for improving the eciency
of the LiO2 cell.
3.5. Air Electrode Architecture

Although electrolyte stability is of paramount importance, air


electrode materials also represent a major technology challenge
in rechargeable aprotic LiO2 cell development. The ultimate
goal of the investigation is to determine how to eectively
increase the specic capacity and power capability of LiO2
cells yet still achieve a long cycle life. Attaining that goal
strongly depends on the materials and their microstructures in
the O2-breathing electrode.62,68,226 Good electronic and ionic
conductivity, fast oxygen diusion, and stable integrity are
required for air electrode materials to achieve high performance
in LiO2 cells. The specic surface area and porosity of the
electrode material are also critical, since the overall capacity and
energy density of LiO2 cells are dependent on the quantity of
the discharge product (Li2O2) that is able to be storeed on or
within the porous air electrode structure. Typically, a larger
surface area provides more surface to uniformly disperse
catalytic particles, if needed, and more active sites to aid the
electrochemical reactions, while a porous structure with an
appropriate pore size provides the space to store the discharge
products. However, further studies revealed that the volume of
the pores in the range of 250 nm rather than the surface area
of the air electrode material appears to be the limiting factor to
determine the discharge capacity of the cell.227,228 For instance,
Cheng and Scott229 showed that Super P carbon with a
relatively low surface area (62 m2/g) demonstrated the highest
discharge capacity among the various carbon materials
investigated, including AC with a very high surface area up to
2100 m2/g. Black et al.23 recently pointed out that, however,
the discharge capacity of a LiO2 cell is not solely determined
by either the surface area or pore size/volume of the air
electrode, but by a more complicated relationship between
these components. The surface area in relation to the types of
pores could compromise the air electrode membrane structure,
while the nature of the carbon material (especially the defects)
may act as an ORR catalyst, both of which will have a signicant
impact on the discharge capacity of the cell. Furthermore, an
ecient catalyst is often needed in the air electrode structure to
facilitate the ORR, and particularly the OER in LiO2 cells,
similar to that in a proton exchange membrane fuel cell. In
some cases, bare carbon-based materials showed some ability to
form Li2O2 when used in conjunction with a relatively stable
electrolyte, since the defects on these carbon-based materials
could serve as the ORR catalyst and the nucleation sites for
Li2O2.61,62
Most of the early studies on catalytic activity toward ORR
and OER of the air electrode materials in aprotic LiO2 cells
were based on the utilization of organic carbonate electrolytes,33,48,51,56,63,75,226,227,230242 which have since been shown
to be unstable to the oxygen reduced species, as discussed
previously. Because the discharge products are now known to
be Li2CO3 or lithium carboxylates, not the desired Li2O2 in the

Figure 16. Optimized high-spin (Li2O2)16 cluster (quintet) structures


from the PBE/PW (top) and B3LYP/6-31G(d) (bottom) levels of
theory. The distances of the short OO bonds and Bader charges (in
parentheses) of OO moieties are shown for the PBE/PW calculation.
The distances are also shown for the B3LYP/6-31G(d) structure along
with the computed Mulliken spins and charges in parentheses. Note
that the fourth unpaired electron is delocalized over other oxygen
atoms. Reprinted from ref 223. Copyright 2012 American Chemical
Society.

important implications for the electrochemistry of reversible


formation of Li2O2 in LiO2 cells, including surface electronic
conductivity and electrolyte surface reactions.
Experimentally, Lu et al.225 recently investigated the
magnetic properties of Li2O2 using electron paramagnetic
resonance (EPR) and superconducting quantum interference
device (SQUID) magnetometry. They showed evidence, for the
rst time, that the Li2O2 formed during the discharge of a cell
with an ether-based electrolyte is paramagnetic, as shown in
Figure 17. The EPR spectra provided direct evidence that the
spin in the discharge Li2O2 product is caused by a superoxidetype structure. The number of spins calculated from magnetic
measurements is also consistent with the predicted one by
DFT. More recent work by Yang et al.174 further conrmed the
presence of a superoxide-like species in the Li2O2 discharge
product by the presence of a peak at 1125 cm1 in the Raman
spectra. The presence of magnetism in the discharge product
(Li2O2) could play an important role in the charge and
discharge chemistries of LiO2 cells, since the superoxide-type
oxygen groups with spin on the Li2O2 surface could enable the
electronic conductivity mechanism that is required for the
reversible formation and decomposition of Li2O2, thus reducing
the charge overpotential. Yang et al.174 observed two distinct
voltage plateaus during charging, at 3.23.5 and 4.24.3 V, in a
LiO2 cell that had petroleum coke-based activated carbon
(AC) as the air electrode material and a TEGDME/LiCF3SO3
electrolyte. The lower plateau corresponds to a form of lithium
O

dx.doi.org/10.1021/cr400573b | Chem. Rev. XXXX, XXX, XXXXXX

Chemical Reviews

Review

Figure 17. Magnetic susceptibility data from SQUID measurements for (a) a pristine carbon air electrode before cycling and a discharged air
electrode with 1NM3/LiCF3SO3 electrolyte and (b) the dierence between the pristine carbon air electrode and the initially discharged air electrode.
(c) Magnetic moment data from SQUID measurement for Li2O2 bulk powder. (d) EPR spectrum of Li2O2 bulk powder at 4 K. The red trace is the
experimental result. The blue trace is the calculated one-line Lorentzian spectrum to t the central peak at the g2 value, and the green trace is the
dierence between the red and blue traces to reveal peaks at g1 and g3. Reprinted with permission from ref 225. Copyright 2013 Wiley-VCH.

performance is strongly believed to depend on the nature of the


catalytic air electrode, in addition to the stability of the
electrolytes. Many catalysts, including metal oxides and
nonprecious and precious metals on a porous carbon support,
have been examined as the air electrode active material for the
oxygen reduction and evolution reactions, showing large
dierences in discharge capacity among dierent catalysts.
Surprisingly, a nearly identical discharge voltage plateau at
about 2.62.7 V is observed for dierent catalysts, similar to
that for the bare carbon without loading of any catalyst. This
eect probably implies that either the ORR kinetics in these
LiO2 air electrodes is limited by the oxygen mass transport
resistance toward the catalysts or carbon itself can provide
sucient ORR activity. Understanding the role of carbon in the
electrochemical reactions in the LiO2 cell could provide
guidance and a baseline for identifying ecient catalysts to
facilitate ORR and OER and thereby improve the cell
performance.
Porous carbon has been chosen as the air electrode material
for almost all the LiO2 batteries investigated so far, mainly
because it can provide sucient charge transfer for the
electrochemical reactions and space for housing the discharge
products. Due to the low mass of the carbon-based air
electrode, it is expected to achieve the highest specic capacity
in LiO2 cells using a metallic lithium electrode. Moreover,
porous carbon often shows certain catalytic activity toward
ORR because of the existence of defect sites on the carbon
surface. Commercial carbon black, in particular, Super P
carbon44,45 and Ketjen Black,48 has been widely used in the
current LiO2 cells. For example, Jung et al.91,136 reported a

cell with the carbonate electrolyte, the claimed catalytic activity,


especially on the charge reaction, appears to be more related to
the electrolyte decomposition, and their true catalytic activity
toward the reversible formation of Li2O2 certainly needs to be
re-evaluated in a more stable electrolyte, such as ether- or
DMSO-based electrolyte. The knowledge gained from the
above research could be benecial for identifying ecient
electrocatalysts for the formation of Li2O2 if a stable electrolyte
can be developed, since the rst step in the formation of either
Li2CO3 or Li2O2 shares the same reaction, i.e., reduction of O2
to O2, yet the charge reaction (i.e., lowering the charge
overpotential) for the LiO2 cell remains a major challenge.
The exact mechanism of the decomposition of Li2CO3 along
with its derivatives in carbonate electrolytes or Li2O2 in ether
electrolyte is up for debate. It can be foreseen that the
decomposition of lithium carbonate and lithium peroxide
during charging of an electrochemical cell will proceed through
a completely dierent pathway because of the dierent
compositions of these compounds. In this sense, the claimed
catalytic activity on charging of a LiO2 cell in a carbonate
electrolyte will not likely apply to that in ether or other
electrolytes where formation of Li2O2 occurs. In the following
sections, we discuss the eect of the catalytic air electrode on
the performance of the LiO2 cell with relatively stable
electrolytes and Li2O2 formation.
3.5.1. Porous Carbon. As mentioned previously, one of the
biggest hurdles for the rechargeable LiO2 battery is the large
overpotential during discharge and, in particular, charge, even at
very low current density (0.010.05 mA/cm2), which results in
low round-trip eciencies and low power capability. This
P

dx.doi.org/10.1021/cr400573b | Chem. Rev. XXXX, XXX, XXXXXX

Chemical Reviews

Review

Figure 18. Evolution of Li2O2 discharge product morphology. Insets show the corresponding discharge voltage prole. (a, b) Galvanostatic discharge
to a capacity of 350 mAh/g at 68 mA/g. Li2O2 particles appear to rst form on the CNT sidewalls as small spheres with <100 nm diameters. (c, d)
Intermediate galvanostatic discharge to 1880 mAh/g at 64 mA/g. Particles appear to develop a toroidal shape as the average particle size increases to
400 nm. (e, f) Full discharge to 7200 mAh/g at 63 mA/g. The discrete particles merge to form a monolithic Li2O2 mass with low porosity. The
capacity and current density are obtained on the basis of the mass of the CNT. Reprinted with permission from ref 61. Copyright 2011 Royal Society
of Chemistry.

LiO2 battery with impressive capacity and rate capability using


Super P carbon as the only air electrode active material in
combination with TEGDME/LiCF3SO3 electrolyte.
Aside from the commercially available carbon black, recent
studies showed that the performance of the new carbon-based
materials could also be very successful when used with a stable
electrolyte. Mitchell et al.61 reported a LiO2 cell using hollow
carbon bers as the air electrode material, which directly grow
on a ceramic porous substrate, eliminating the need for a
binder. These all-carbon-ber electrodes demonstrated an
exceptionally high discharge capacity of 7200 mAh/gcarbon,
translating to a gravimetric energy of 2500 Wh/kgcarbon, which
is 4 times higher than that of the state-of-the-art lithium
intercalation compounds such as LiCoO2 (600 Wh/kg). They
attributed such high capacity of the cell to low carbon packing
and highly ecient utilization of the available carbon mass and
void volume for Li2O2 formation. With the unique structure of
the carbon ber, they were able to monitor the Li2O2 formation
and morphological evolution during discharge as shown in
Figure 18, where the toroid-shape Li2O2 was rst reported. The
visualization of Li2O2 morphologies upon discharge and
disappearance upon charge represents a critical step toward
understanding key processes that limit the rate capability and
low round-trip eciencies of LiO2 batteries. In a follow-up
paper by the same group,101 a more detail mechanism of
morphological evolution of Li2O2 particles during electrochemical growth was reported using a freestanding carbon
nanotube (CNT) air electrode in a DME-based electrolyte. The
morphology of Li2O2 at the rst discharge cycle appears to be a
function of both the discharge rate and capacity (see Figure
19), which is related to the nucleation and growth process of
Li2O2 on CNT. At low discharge current density, sparse
particles rst nucleate on the CNT sidewalls as stacked thin
plates, which spontaneously splay so that secondary nucleation
of new plates eventually leads to the development of disk- and
toroid-shaped particles, as shown in Figure 19a,b. The highresolution TEM images on CNT suggest an instantaneous
nucleation of Li2O2 at relatively few sites early in the discharge
process. Thin plates of Li2O2 with the large facet of the [001]

Figure 19. SEM and TEM micrographs of Li2O2 particles on


electrodes consisting of freestanding CNT carpets discharged at low
and intermediate gravimetric rates. (a) SEM and (b) TEM
micrographs of an electrode discharged at 10 mA/gcarbon to 200 mA
h/gcarbon, with disk-shaped particles and bare CNT sidewalls. (c) SEM
and (d) TEM micrographs of an electrode discharged at 90 mA/gcarbon
to 13 000 mAh/gcarbon, with a high density of disk particles and a thin
coating of discharge product present on the sidewalls of the CNTs.
Insets: Higher magnication TEM images of the CNT sidewalls,
indicated by a dashed yellow line, showing the (b) absence and (d)
presence of small particles (scale bar = 20 nm). Reprinted from ref
101. Copyright 2013 American Chemical Society.

plane were found in both disk- and toroid-like particles,


consistent with theoretical calculations of the equilibrium Wul
shape for Li2O2 and a layer-by-layer growth mode. In contrast,
at higher discharge rates, copious nucleation of equiaxed Li2O2
Q

dx.doi.org/10.1021/cr400573b | Chem. Rev. XXXX, XXX, XXXXXX

Chemical Reviews

Review

Through the above research work, it is clear that the Li2O2


morphology (e.g., shape and characteristic thickness) and
structure (e.g., crystallinity and surface vs bulk composition) are
important parameters that can inuence the discharge capacity,
rate capability, and cyclability of a LiO2 cell. Therefore, it will
be a major contribution to the development of LiO2 cells if
the parameters associated with the Li2O2 morphology and
structure can be nely tuned. Understanding the link between
the morphology and structure of Li2O2 and the electrochemistry will be critical for identication of new approaches to
reduce the overpotentials on discharge and, more signicantly,
on charge.27
3.5.2. Catalysts on Porous Carbon. Since the rst
demonstration of a lithiumoxygen system using pyrolyzed
cobalt phthalocyanine over carbon as the air electrode,10 many
new air electrode materials have been reported, particularly in
the area of new catalysts for improving the battery
eciency.28,30,32,59,244 As discussed in section 3.3, most of the
early studies on these catalytic air electrode materials were
based on organic carbonate electrolytes, which were later
proved to catalyze the formation/decomposition of Li2CO3
and/or the decomposition of the carbonate electrolyte rather
than the formation/decomposition of Li2O2. These catalysts
include transition-metal oxides, nonprecious metals, and
precious metals (and alloys), as reviewed by Shao et al.,42
and their true catalytic activity in LiO2 systems needs to be reevaluated. Below, we only discuss some recent research work
on catalytic air electrodes that form and decompose Li2O2, i.e.,
in more stable electrolytes.
To investigate the electrocatalytic activity of any catalyst on
reaction 4 (Figure 3) in an aprotic LiO2 cell, the formation of
Li2O2 is a prerequisite. As discussed in section 3, the choice of
the organic electrolyte is the most critical factor determining
whether Li2O2 will form during cell discharge. Oftentimes, air
electrode materials, including electrocatalysts, inuence the
discharge products of the LiO2 cell, in particular the
formation of Li2O2. For example, Freunberger et al.83
investigated three carbons, namely, Super P, Ketjen Black,
and Black Pearls, in LiO2 cells with an ether-based electrolyte,
and they identied Li2O2 as the dominating discharge product.
However, the electrolyte also decomposed, as was evident in
their FT-IR results (through (CO) bands), which indicated
the presence of polyethers/esters on the air electrode surface.
They also found a much higher concentration of polyethers/
esters in Black Pearls than in Super P or Ketjen Black, probably
indicating a higher catalytic activity of Black Pearls toward the
decomposition of the electrolyte. A similar eect on the air
electrode was also reported by Allen et al.153 when they
compared the cyclabilities of a Au electrode and a glassy carbon
electrode. Nevertheless, the catalytic composition of the air
electrode material does signicantly impact the discharge
products and, therefore, the cyclability of the LiO2 cell.
In general, the electrocatalyst is critical to improve the rate
capability, cyclability, and round-trip eciency of the LiO2
cell. Earlier studies demonstrated that the catalysts investigated
to date probably did not change the mechanism of the
discharge reaction toward the formation of Li2O2, but they may
decrease the activation energy for oxygen reduction, which
translates into a slightly higher discharge voltage in some cases.
Various catalysts have been examined to lower the activation
barriers, such as metal oxide, precious metals, and nonprecious
metals, the mostly widely studied of which is MnO2.45,216,245
Since MnO2 nanoparticles can exhibit a large variety of

particles precedes growth of disks and toroids. A much higher


density of Li2O2 particles with nonuniform shapes was
observed, and the CNT sidewalls appeared to be coated by
numerous small particles (Figure 19c,d). Their study also
demonstrated that the coating of the carbon surface with small
irregular particles precedes growth of well-dened larger diskand toroid-shaped particles with increasing discharge rates.
Findings on the toroidal morphology of Li2O2 are interesting
and critical, since it leads to very nonuniform surface coverage,
which is benecial to increasing the discharge capacity of the
cell due to easier oxygen diusion at the late stage of the
discharge. Therefore, understanding and controlling the
nucleation and morphological evolution of Li2O2 particles
upon discharge is the key factor to achieving high volumetric
energy density of the LiO2 cell. Several research groups have
also observed the formation of Li2O2 toroid-shaped particles on
dierent carbon air electrodes with or without additional
catalysts.44,174,175,243
Graphene-based materials have also been investigated as the
air electrode due to their light weight and conductive and
catalytically active surface. For instance, Xiao et al.62 reported a
primary Liair battery containing hierarchically porous
functionalized graphene sheets (FGSs) delivers an extremely
high capacity of 15 000 mAh/gcarbon, demonstrating the
potential application of graphene in the LiO2 system. Such
signicantly improved performance of the graphene-based air
electrode is attributed to its unique morphology and dierent
Li2O2 deposition mechanism. The FGS electrode has numerous
large tunnels facilitating continuous oxygen ow into the air
electrode, while other small pores provide ideal triphase
regions for the oxygen reduction, as shown in Figure 20a,b.

Figure 20. Morphologies of the graphene-based air electrode. (a, b)


SEM images of as-prepared FGS (C/O = 14) air electrodes at dierent
magnications. (c, d) Discharged air electrode using FGSs with C/O =
14 and C/O = 100, respectively. Reprinted from ref 62. Copyright
2011 American Chemical Society.

Unlike the toroid-shaped particles reported by Mitchel et al.61


as discussed above, the deposited Li2O2 forms isolated
nanosized islands on FGSs (Figure 20c,d), supported by
their DFT calculation which suggests that the aggregation of
Li2O2 clusters is energetically unfavorable in the vicinity of
defective sites of FGSs.
R

dx.doi.org/10.1021/cr400573b | Chem. Rev. XXXX, XXX, XXXXXX

Chemical Reviews

Review

crystallographic structures (e.g., , , , , and forms) and


morphologies, controlled synthesis of nanostructured MnO2
with high crystalline purity and uniform morphology is critical
to achieving precisely tailored properties and, thus, better
performance in their applications in LiO2 batteries. To
address this issue, Truong et al.125 produced high-quality
nanosheet-based -MnO2 microowers, -MnO2 nanowires,
and -MnO2 nanotubes in large quantity through the
microwave-assisted hydrothermal reduction of potassium
permanganate in the presence of hydrochloric acid, as shown
in Figure 21. They found that the chemical reaction determines

oxygen reduction and evolution reactions and, therefore, lead to


signicant enhancement of the electrochemical performance of
LiO2 cells. In this case, a low charge overpotential (3.5 V) was
achieved by using the as-prepared MnO2/C composite as the
air electrode material with a TEGDME-based electrolyte.
Using a dimethoxyethane-based electrolyte, McCloskey et
al.88 investigated the electrocatalytic role of Au, Pt, and MnO2
nanoparticles and found none performed better than carbon.
They found the charge potential at the beginning of the charge
cycle to be low (3.4 V), but it rose to above 4 V during the
rst cycle. Black et al.175 studied the role of catalysts and
peroxide oxidation in LiO 2 batteries to address the
aforementioned question on the ecacy or necessity of
catalysis in the aprotic system. They demonstrated that charge
overpotential could be signicantly reduced when a nanocrystalline Co3O4/RGO (reduced graphene oxide) composite is
included as part of a carbon-based air electrode, as shown in
Figure 22. XRD results conrmed the presence of Li2O2 on the
rst discharge, and SEM images showed that these Li2O2
particles have a toroid shape. Black et al. indicated that
Co3O4/RGO clearly enhances the kinetics of mass (or surface)
transport for both oxygen reduction and evolution rather than
conventional electrocatalytic eects of the particles. However,
the signicant improvement on cycling appears to be

Figure 21. SEM images of MnO2 products obtained at dierent


reaction times: (A) 2, (B) 5, (C) 15, (D) 30, (E) 120, and (F) 360
min. The temperature of the reaction solutions and the atmosphere
pressure above the reaction solutions were 150 C and 70 psi,
respectively. The concentrations of KMnO4 and HCl in the reaction
solution were 0.05 and 0.2 M, respectively, before the reaction was
initiated. The insets are the high-magnication SEM images of the
corresponding structures. The scale bars in (F) also apply to (A)(E).
Reprinted from ref 125. Copyright 2013 American Chemical Society.

the formation of -MnO2 microowers in the earlier stages of


the process, while the two-step Ostwald ripening process is
responsible for the crystalline and morphological transition
from the -MnO2 nanosheets to -MnO2 nanowires and
nanotubes in the late stage of the reactions. In terms of the
electrocatalytic activity for these dierent MnO2 nanoparticles,
the single-crystal -MnO2 nanotubes exhibit much better
performance to catalyze the electrochemical processes involved
in aprotic LiO2 cells using TEGDME-based electrolytes,
leading to a signicant improvement in specic capacity and
cycle life. Qin et al.45 developed an in situ wet-chemistry
approach to incorporate MnO2 nanorods into porous carbon.
This process avoids any aggressive post-treatment of the carbon
materials and, thus, preserves the original structure of carbon.
As-prepared MnO2/C composites with porous structures and
high specic surface area provide more active sites for enabling

Figure 22. (a) First dischargecharge prole for LiO2 cells with
Ketjen Black (KB) or Co3O4/RGO/KB at 25 C at a current rate of
140 mA/g. (b) XRD patterns of the cells in (a) on rst discharge.
Reections of Li2O2 are marked. (c) SEM micrographs of the
discharged electrodes at 6000 mAh/g: (i) Co3O4/RGO/KB and (ii)
KB. (d) Chronoamperometry showing normalized current evolution
with time at 2.25 V. (e) Linear sweep voltammetry following a hold at
2.25 V for 1 h. (f) Voltage prole on charge for cells containing
chemically deposited Li2O2 in the presence of either Co3O4/RGO/KB
or KB. The specic capacity and current density were based on the
mass of the active materials. Reprinted with permission from ref 175.
Copyright 2013 Wiley-VCH.
S

dx.doi.org/10.1021/cr400573b | Chem. Rev. XXXX, XXX, XXXXXX

Chemical Reviews

Review

Figure 23. (a) Voltage prole during dischargecharge of cells (to 1000 mAh/g) based on Super P carbon, Super P carbon coated with Al2O3, and
Al2O3-coated Super P carbon further coated with Pd nanoparticles. The electrolyte used is TEGDME/LiCF3SO3. (b) Voltage prole during
dischargecharge of cells (to 500 mAh/g) based on Al2O3-coated Super P carbon further coated with Pd nanoparticles. The electrolyte used is
TEGDME/LiCF3SO3. The specic capacity and current density were based on the mass of the active materials. Reprinted with permission from ref
248. Copyright 2013 Nature Publishing Group.

overwhelmed by the deposition of side products resulting from


parasitic reactions with the electrolyte. This eect was
attributed to the high reactivity of liberated *O2 and/or
peroxide with the TEGDME electrolyte on charge. If this
nding is applicable to a range of metal oxide catalysts and
electrolytes, it could provide guidance to better understanding
surface reactivity and tailoring of electrodes and electrolytes to
overcome the limitations.
To achieve a true LiO2 battery, the charge (oxygen
evolution reaction) seems to be more critical than the discharge
process, because of the presence of large charge overpotentials.
In addition to the lack of an eective electrocatalyst, a number
of other factors may contribute to the large charge overpotentials observed in the LiO2 cells. For example, poor
electronic conductivity of the Li2O2 discharge product may
limit the charge process, which depends on electron transport
to the Li2O2/electrolyte interface.221,224,246 The large charge
overpotentials could also result from the contaminants in the
Li2O2 discharge product from the decomposition of an
electrolyte constituent, such as Li2CO3. Indeed, McCloskey et
al.247 showed that decomposition of the electrolyte may occur
at defect sites on the carbon surface, while Bruce and coworkers145 demonstrated that oxidation of the carbon surface
can take place at voltages as low as 3.5 V during charge in the
presence of Li2O2. Lu et al.248 recently showed promising
results for tackling the charge overpotential problem by a
rational design of new air electrode architecture for LiO2 cells.
The air electrode architecture addresses the electrolyte
decomposition problem by use of a porous carbon passivated
by a protective Al2O3 coating. Palladium nanoparticles attached
to the surface by atomic layer deposition act as eective
electrocatalysts, which promote the growth of a nanocrystalline
form of Li2O2 having the electronic transport properties needed
to lower the charge potential. The resulting charge overpotential is only about 0.2 V, as shown in Figure 23. This new
air electrode architecture opens new avenues for LiO2 battery
development through optimization of the metal nanoparticle
and the carbon support, as well as incorporation of new
protective coatings on air electrodes.
Recently, using a liquid-crystal template and subsequent
oxidation by a chemical agent, Oh et al.243 reported a novel
mesoporous (MP) metallic oxide as part of the air electrode
material for an aprotic LiO2 cell. Air electrodes of this
material show a high reversible capacity of 10 000 mAh/gMP

(1000 mAh/g with respect to the total electrode mass


including the peroxide product). The test results indicated that
the catalytic activity of this MP oxide toward the reversible
oxygen reduction and evolution in the LiO2 cell is surprisingly
high, and its reversible capacity surpasses that of any
nanocrystalline oxide reported so far. This MP oxide also
showed a lower charge potential for oxygen evolution from
Li2O2 (by 0.5 V) than pure carbon. This enhanced
electrochemical performance is attributed to the high fraction
of surface active sites in the MP metallic oxide and its unique
morphology and variable oxygen stoichiometry. This strategy
for creating porous metallic oxides may pave the way to new air
electrode architectures for the reversible aprotic LiO2 cell.
3.6. Lithium Electrode

Metallic lithium is the current choice of the metal electrode


material for LiO2 batteries, which is expected to achieve the
highest energy density at the cell level since lithium itself has an
extremely high specic energy (3860 mAh/g) and a low
negative potential (3.04 V vs standard hydrogen electrode,
SHE). In rechargeable Li ion batteries, however, utilization of
lithium as the metal electrode has been a long-standing
problem remaining to be solved. The main issue is dendrite
formation on the lithium side during repeated charge/discharge
cycles, which can cause internal short circuiting of the Li ion
cell and, thus, a severe safety concern.249,250 Low Coulombic
eciency is another issue facing the lithium electrode. It has
been proposed that the continuous growth of the SEI on the
lithium electrode surface under uneven current distributions
and formation of irreversible dead lithium are responsible for
the dendrite formation.249 Despite decades of extensive
research, a lithium metal electrode has yet to be deployed in
any secondary lithium batteries with a long cycle life. A number
of research papers on the topic of Li ion batteries, and even
some recent reviews on LiO2 batteries,32,37,42 have addressed
the dendrite formation problem; therefore, we do not discuss
this issue in further detail here. Moreover, dendrite formation
has not been observed in the LiO2 cells tested to date. In fact,
the reactions occurring on the interface between the Li
electrode and electrolyte in LiO2 cells are far more
complicated than expected due to O2 crossover from the air
electrode. For this reason, we only discuss the eect of oxygen
crossover on the degradation of the Li electrode and the
T

dx.doi.org/10.1021/cr400573b | Chem. Rev. XXXX, XXX, XXXXXX

Chemical Reviews

Review

Figure 24. Proposed mechanism and energetics for the formation of LiOH from an ether-based solvent (TEGDME) at the electrode upon reduction
and oxygen crossover. Also shown in the box is a mechanism for the further formation of lithium alkyl carbonates and carbonates from the resulting
aldehydes. All energetics are changes in enthalpies (eV) in solution ( = 21) at 298 K computed at the B3LYP/6-31+G(d) level of theory Reprinted
with permission from ref 57. Copyright 2013 Wiley-VCH.

TEGDME anion radical (from the electrochemical reduction at


the Li electrode) is exothermic. These reaction pathways will
nally lead to the formation of the LiOH and lithium
carbonates. As the concentration of these species increases,
the insoluble LiOH and Li2CO3 will deposit on the Li
electrode. Assary et al. have also carried out a detailed
characterization study of the Li electrode during various stages
of discharge and charge using in situ XRD and FTIR
measurements, which clearly indicate the formation of LiOH
and Li2CO3 on cycling. As a consequence, the Li+ ion migration
will be signicantly blocked by these insoluble products, leading
to poor cell performance. This study indicated that controlling
SEI reactions at the lithium electrode through suitable
membranes or passivation lms is essential for achieving
good performance of LiO2 cells. Previous studies also
suggested the need for an ecient protective layer for metallic
lithium to avoid decomposition of the Li electrode due to
contamination by discharge/charge products in the LiO2
cells.37,60,251253
Recently, Shui et al.108 investigated the reversibility of the
anodic lithium inside an operating LiO2 cell using spatially
and temporally resolved synchrotron XRD and a 3-dimensional
microtomography technique. The combined electrochemical
and in situ XRD results revealed a continuous accumulation of
LiOH on the lithium electrode under both charge and
discharge conditions, along with partial recovery of lithium
during the charge cycle, as shown in Figure 25. Since LiOH is

possible solutions to improve the Li electrode performance in


LiO2 cells.
As discussed earlier, the current major focus of studies on
aprotic LiO2 batteries is the roles that electrolyte decomposition and the catalytic oxygen electrode play in cell
performance. It is well established that electrolyte decomposition in LiO2 systems mainly arises from chemical or
electrochemical reactions at the electrolyteair electrode
interface. The electrolyte decomposition at the electrolyte
metal electrode interface, if any, has not been investigated in
much detail yet. In a Li ion cell, Aurbach et al. have shown that
lithium metal electrodes fail when carbonate solvent (e.g.,
ethylene carbonate) is reduced and decomposes to form surface
layers of ROCO2Li, ROLi, and Li2CO3 species.249 Since ethers
have reduction potentials similar to those of carbonates, it is of
interest to investigate the possible decomposition mechanism at
the lithium electrode, especially in the presence of O2 crossing
over from the air electrode and the possible eects on the
performance of the LiO2 cell.
Assary et al.57 investigated the stability of an ether-based
electrolyte (TEGDME/LiCF3SO3) at the lithium electrode in a
LiO2 cell where oxygen crossover occurs. In their study, highenergy XRD, FTIR spectroscopy, and DFT were used to
determine the possible reactions that may occur under an O2
environment. The DFT calculations indicated that the presence
of O2 at the Li electrode will favor the reaction pathways shown
in Figure 24, considering that the binding of O2 with the
U

dx.doi.org/10.1021/cr400573b | Chem. Rev. XXXX, XXX, XXXXXX

Chemical Reviews

Review

Figure 25. (a) Dischargecharge voltage prole of the operando LiO2 cell as a function of the cycling time. The stages at which XRD data sets
were collected are marked as i, ii, iii, ... on the curve. (b) Four representative XRD sets taken at cycling stages i, iii, vi, and ix; each set consisted of
seven selected XRD patterns collected at various electrodeseparator interfacial depths (marked as 1, 2, ..., 7 of dierent colors on the right of each
set). (c) Expanded XRD data set at stage ix covering the whole electrode through the adjacent separator region. The inset is a representative XRD
pattern taken from the position marked by a blue line that includes references of Li and LiOH. Reprinted with permission from ref 108. Copyright
2013 Nature Publishing Group.

neither a Li+ ion nor an electron conductor, a continuous


supply of Li+ from the metallic lithium electrode would be
problematic if LiOH formed on the lithium surface as a dense
lm. Using a microtomography technique, Shui et al.
demonstrated that a large number of microscopic tunnels
exist within the LiOH layer. These tunnels provide a pathway
for sustained ion transport, since they enable the connection of
the metallic lithium with the electrolyte. Therefore, the battery
can be operated until the total consumption of lithium has
occurred, assuming that the air electrode and electrolyte can
survive for that long period.
Because the separators currently used in Liair cells cannot
prevent O2/H2O diusion to the Li electrode, oxygen crossover
can be a signicant problem, leading to fast decay of the Li
electrode, as discussed above. It is critical to develop thin, active
membranes that can be embedded within passive porous
polymeric membranes to eectively eliminate the O2 crossover
to the Li electrode, which is responsible for limiting the
reversibility of LiO2 cells. To address this issue, Truong et
al.254 fabricated single-crystal silicon membranes, which are
capable of conducting Li+ ions through their lattices to support
current density as high as 1 mA/cm2. However, compared with
the best NASICON-type lithium ion conducting membranes,
the Li+ conductivity of the single-crystal Si membranes is still
too low (34 orders lower). Their ionic conductivity might be
improved by doping the surface lattices of single-crystal Si
membranes, lowering the interfacial energy barriers for Li+
insertion/extraction from the Si lattices, or partially lithiating
the single-crystal Si membranes to increase the concentration of
Li+ ions in the membranes and thereby reduce their diusive
resistance. In addition, direct contact of the single-crystal Si
membranes with Li metal should be avoided due to the severe
electrochemical reactions, although silicon is stable in many
electrolytes, including neutral/acidic aqueous electrolytes and
aprotic organic electrolytes.

4. AQUEOUS LIO2 BATTERIES


As presented above, we extensively reviewed the challenges and
advances in the aprotic LiO2 system, which is garnering the
major fraction of the current research on Liair batteries.
Unlike the aprotic electrolyte, for which numerous types of
liquid or solid ionic conductors can be selected, the aqueous
electrolyte is limited to acidic or basic solutions only. In the
aqueous Liair battery, electrolyte solvent, e.g., H2O, is not a
limiting factor in the cell performance, which is its main
advantage over the aprotic system. In addition, incombustible
aqueous electrolytes circumvent the safety issue,255 which is a
major concern for the organic electrolytes in an open cell
conguration. However, due to the dierent electrochemical
reactions involving Li and O2, the gravimetric and volumetric
capacities of an aqueous LiO2 cell are much lower compared
to those of an aprotic cell.50
A typical aqueous Liair cell conguration is shown in
Figure 2b. During the cell discharge, oxygen is reduced on the
air electrode via the ORR, similar to that in the aprotic system.
However, water in the electrolyte solution also participates in
the electrochemical reactions for the aqueous Liair system.
On charging of the secondary aqueous Liair cell, the OER
takes place on the air electrode to produce Li+, O2, and
electrons. Catalysts are often used to promote the 4e ORR
and OER with increasing activities in secondary cells. To
support such catalysts, porous carbon along with a possible
binder has been widely employed as part of the air electrode
material, which also provides electronic conduction. In some
cases, a gas diusion layer (GDL) is attached to the active air
electrode to allow rapid and uniform diusion of oxygen to the
active reaction sites. For a rechargeable aqueous Liair cell, the
GDL must have both hydrophobic and hydrophilic pores to
supply O2 during ORR and transport water from the air
electrode to stop ooding, respectively. A Li+ conducting lm
must be present on the lithium electrode to prevent an
excessively vigorous reaction between lithium metal and H2O.
V

dx.doi.org/10.1021/cr400573b | Chem. Rev. XXXX, XXX, XXXXXX

Chemical Reviews

Review

including Li2O/Al2O3/TiO2/P2O5 and Li2O/Al2O3/Ge2O/


P2O5. However, while lithium metal is not compatible with
LiC-GC, that problem could be overcome by introducing a Li+
conducting interlayer between the Li electrode and LiC-GC
membrane. More details on such an interlayer can be found in a
recent review by Crowthe and Salomon.265
It is well established that catalysts are required to improve
the ORR and OER in conventional fuel cells. Since the air
electrodes of the primary and secondary aqueous Liair cells
operate in a fashion similar to that of the polymer electrolyte
fuel cell and unitized regenerative fuel cell, respectively, the
catalysts developed for those fuel cells could serve as
benchmark catalysts for aqueous Liair cells. As a proof-ofconcept, Pt is widely used in most initial studies on the
aqueous Liair battery as the electrocatalyst for ORR due to its
catalytic activity being dicult to match, particularly in acidic
media.251,266268 Since Pt is relatively rare and expensive, care
must be taken to minimize the Pt loading while still
maintaining high electroactivity. In an attempt to explore
alternatives to the Pt catalyst, Zhou and co-workers54,238,255,269,270 recently investigated the electroactivity of
transition metals and metal oxides in the aqueous Liair
system. Air electrodes fabricated with Mn3O4/activated
carbon54 as a catalytic layer coupled with a GDL delivered a
specic capacity of 50 000 mAh/g based on the total mass of
the catalytic electrode (carbon + binder + catalyst), a result
indicating that Mn3O4 appears to have high ORR catalytic
activity in basic solution. This is the highest capacity reported
so far for any LiO2 cells in aprotic or aqueous systems.
In general, monofunctional catalysts only increase the activity
of one particular reaction, either the ORR or the OER. For
example, Pt is a model ORR catalyst as mentioned earlier. Also,
IrO2 is known to improve the rate of OER but does not
improve the rate of ORR. For secondary aqueous Liair cells, a
bifunctional catalyst with the capability to catalyze both ORR
and OER is mandatory. Wang et al.271 studied CoMn2O4
nanoparticles on graphene as a catalyst for a secondary Li
air cell in alkaline aqueous electrolyte, which shows a high ORR
activity, comparable to that of a Pt/C catalyst, and a
substantially improved OER activity. They also found that
CoMn2O4 nanoparticles alone are responsible for the OER with
graphene as an inert support and electronic conductor. A recent
study on perovskite catalyst272 for ORR and OER in aqueous
solution provided further fundamental guidelines to search for
more eective catalysts used in the aqueous Liair batteries.

In general, the discharge products are soluble in the aqueous


solution, and the discharge potentials of the aqueous Liair
cells are slightly higher than those of the aprotic systems,
normally in the range of 3.03.3 V vs Li/Li+, depending on the
pH of the aqueous electrolyte. The typical aqueous electrolyte
is a mixture of LiOHLiCl in water,52,256 where the
concentration of LiCl and/or pH of the solution can vary.
Depending on the pH of the electrolyte, several electrochemical
reactions in the aqueous Liair system could occur at the air
electrode. In a basic aqueous solution, O2 reduction on the air
electrode involves H2O as a reactant, as shown in the following
reaction:
2Li + 1 2 O2 + H 2O 2LiOH

(4.1)

The discharge product from this reaction is soluble LiOH,


which has limited solubility in water at room temperature (12.8
g/100 g of H2O). The concentration of OH rises with
increasing depth of discharge until it reaches the solubility limit
of LiOH, which will then precipitate out of the solution as a
monohydrate, LiOHH2O.257 Such excess precipitates could
clog the pores of the air electrode and, therefore, block the
further O2 diusion, if they are not removed mechanically.
Considering the standard equilibrium potential, the cell
discharge voltage can be slightly higher for an aqueous Liair
battery in an acidic electrolyte, which also leads to dierent
discharge products. For instance, in a mildly acidic medium, the
reaction of Li with O2 results in the formation of LiCl,258 as
shown in the following reaction:
2Li + 1 2 O2 + 2NH4Cl 2LiCl + 2NH3 + H 2O
(4.2)

Another example of a reaction in an acidic solution is the


formation of Li2SO4 in the case in which H2SO4 is used as part
of the aqueous electrolyte.258 The main advantage of operating
the Liair cell in an acidic solution is that the formation of
insoluble carbonates, as in basic solution, can be avoided, yet
acidic (or neutral) solution will eventually turn into alkaline
solution during the long-term cycling of a cell. More
importantly, it is very challenging to develop an eective
electrocatalyst for oxygen reduction and evolution reactions in
acidic electrolytes; therefore, for the most part, selection of the
electrocatalyst is limited to the expensive noble metals. In
addition, the specic energy of an aqueous Liair system in
acidic electrolyte is much lower than that in basic electrolyte.
The biggest challenge in realizing an aqueous Liair battery,
primary or secondary, is to prevent the corrosion of metallic
lithium by H2O, which is still a major and continuing eort for
research and development nowadays. As mentioned earlier,
preventing H2O from reaching the Li metal electrode requires a
barrier layer that allows the transport of Li+ only between the
lithium electrode and aqueous electrolyte. A barrier using a
thick passive KOH lm12,13 on the lithium electrode or a
hydrophobic polymer membrane between the Li electrode and
electrolyte259 proved to be unsuccessful since water transport
through these layers could not be completely avoided. In 2004,
the PolyPlus Battery Co.260,261 developed a Li ion conducting
glass ceramic (LiC-GC) to separate the lithium metal from the
electrolyte, which allows the safe operation of the aqueous Li
air battery by blocking H2O from reaching the lithium
electrode. Further developed by Fu and patented by Ohara
Corp.,262264 the most successful example of such LiC-GCs to
date is the single Li+ conducting NASICON-type glass ceramic,

5. CONCLUDING REMARKS AND PERSPECTIVES


Rechargeable LiO2 battery technology is still at the infant
stage, although signicant progress has been achieved, as we
have presented in this review. We mainly reviewed the research
activities and achievements of the aprotic LiO2 battery, since
it has dominated the research eorts for current LiO2 systems.
We have discussed both experimental and modeling results
with regard to the critical factors that have a signicant impact
on the sustainable rechargeability of aprotic LiO2 cells,
including the stability of the electrolyte and its inuence on the
electrochemical performance of the cell, the structure and
magnetic properties of Li2O2 and their relevance to the LiO2
battery, the design and optimization of the air electrode
structure with dierent types of catalysts, and the eect of
oxygen crossover on the Li electrode. Without a doubt,
substantial challenges exist for each component of the aprotic
W

dx.doi.org/10.1021/cr400573b | Chem. Rev. XXXX, XXX, XXXXXX

Chemical Reviews

Review

including FT-IR, 51 XPS, 45,171 Raman, 84,275 NMR, 206


DEMS,59,78 and electron microscopy,136,248 becomes important.
Study of the cell during discharge and charge using in situ highenergy X-ray techniques is expected to provide insight into the
structural changes at the atomic level.276 X-ray absorption ne
structure spectroscopy (XAFS) is another powerful tool for
studies of the electronic and atomic structures of the catalysts in
action,216 due to the penetrating ability of hard X-rays, which
also allows in situ characterization in a suitably designed cell or
reactor. An important advantage of the XAFS technique is the
element-specic nature, which permits investigation of the local
structure of a particular constituent element in a composite
catalyst. In addition, nonresonant inelastic X-ray scattering
(NIXS) using hard X-rays can provide bulk sensitive
information on Liair battery discharged products with
element specicity.277 NIXS, also known as X-ray Raman
scattering, can be considered as an alternative approach toward
obtaining soft X-ray absorption spectroscopy like information
by using hard X-rays, despite crystalline or amorphous
components being present. Theory and computational
modeling55,278 have already played an integral part in the
research and development of electrolytes and catalysts for the
LiO2 battery and should be continued for predicting the
stability of the electrolyte and for screening potential catalyst
candidates for optimal LiO bond breaking and making.
In the case of the aqueous Liair system, a better
understanding of LiO2 electrocatalysis is required at the
current stage, since the LiO2 electrochemistry is unique and
dierent from that of conventional electrocatalysis. The
successful development of any aqueous Liair batteries severely
relies on the prevention of direct contact of the lithium metal
electrode with water. The most innovative approach to address
this issue is the introduction of Li ion conducting glass
ceramics. However, these ceramics are generally fragile and
highly resistive at low temperature. Moreover, they may not be
very stable in strong acidic or basic media. Future research and
development of large and more exible LiC-GC membranes
will be greatly benecial to the aqueous LiO2 system.
Searching for eective catalysts, in particular, with respect to
OER, will be a key challenge for rechargeable aqueous Liair
cells.

LiO2 cells. We have also briey reviewed the recent research


progress on the aqueous Liair system.
Organic electrolytes (both solvents and lithium salts) play
the most critical role in development of an aprotic LiO2 cell
and determine whether a truly rechargeable LiO2 battery can
be realized. One of the main challenges (probably the greatest)
at the current stage for aprotic LiO2 cells is the search for
stable electrolytes. Carbonate-based electrolytes have been
widely used in most of the initial research work, but it is now
universally recognized that these electrolytes decompose in the
presence of the superoxide radicals. Nevertheless, despite the
severe instability of these electrolytes, many research projects
are using them to investigate the catalytic activities of the air
electrode materials. The catalytic activity of the air electrode
materials needs to be re-examined in more stable electrolytes.
Ether-based electrolytes seem to be relatively stable in the
presence of the reduced oxygen species; however, their stability
during charge, especially at high voltage, remains unclear.
Lithium salt, an indispensable component of the electrolyte,
deserves much more attention since it may have a positive
eect on the electrolytes stability in aprotic LiO2 cells.
Without question, searching for a fully stable electrolyte in the
oxygen-rich electrochemical environment is the research
priority at present. Design of a robust strategy for eectively
screening the stability of various electrolytes would be greatly
benecial to the development of a LiO2 battery for practical
application.
In parallel with the importance of the electrolyte,
investigation on how the porous air electrode architecture
aects the formation of the discharge product, Li2O2, and the
specic capacity of the cell is still of great interest. It is
necessary to understand the key limiting factors to determine
the capacity, rate capability, and cycling eciency of aprotic
LiO2 cells. Porous carbon with or without additional catalyst
is the current choice of the air electrode material, although a
few noncarbon air electrode materials have been reported.59,147,273 However, the mechanism of Li2O2 growth on
the porous air electrode during cell discharge and the
subsequent decomposition of Li2O2 on charge is still debatable;
this matter needs to be further claried to develop more
ecient catalysts for the aprotic LiO2 cell.
The lithium electrode has been a historic problem in any of
the Li battery systems, while the long-term cycling of the Li
electrode has yet to be demonstrated. In terms of protecting the
Li electrode from oxygen crossover, controlling reactions of the
electrolyte at the Li electrode through suitable membranes or
passivation lms will be essential for achieving good performance with aprotic LiO2 cells. These membranes should meet
the following criteria: (1) block diusion of oxygen from the air
electrode to the lithium electrode, (2) allow the transport of Li+
to support current ow, and (3) exhibit excellent mechanical
exibility and stability to be compatible with the mechanical
exibility of the supporting polymer membranes and battery
design/processing. Engineering active membranes with a
nanometer-scale thickness could potentially meet these
criteria.274
To study the electrolyte stability and electrocatalytic process
in the aprotic LiO2 system, advanced research tools from both
experimental and theoretical modeling are necessary. Using a
single characterization technique to identify the possible
reaction products in the cell sometimes leads to misleading
results, simply due to the complexity of these compounds. In
this sense, the integration of multiple characterization tools,

AUTHOR INFORMATION
Corresponding Authors

*E-mail: yksun@hanyang.ac.kr.
*E-mail: wufeng863@bit.edu.cn.
*E-mail: amine@anl.gov.
Notes

The authors declare no competing nancial interest.


X

dx.doi.org/10.1021/cr400573b | Chem. Rev. XXXX, XXX, XXXXXX

Chemical Reviews

Review

Biographies

Jin-Bum Park was born in 1985 in Daegu, Korea. In 2011, he received


a Bachelor of Science degree in chemical engineering from Hanyang
University. He is presently a Ph.D. candidate in the Department of
Energy Engineering at Hanyang University, Korea, under the
supervision of Professor Yang-Kook Sun. His research focuses on
materials development in the elds of energy conversion and storage,
such as air electrodes and electrolyte materials for Liair batteries.

Jun Lu is a DOE-EERE postdoctoral fellow in the Vehicles


Technology Program. His research interests focus on electrochemical
energy storage and conversion technology, with the main focus on
beyond Li ion battery technology. Dr. Lu earned his bachelors degree
in chemistry physics from the University of Science and Technology of
China (USTC) in 2000. He completed his Ph.D. in the Department of
Metallurgical Engineering at the University of Utah in 2009, with his
major research on metal hydrides for reversible hydrogen storage
application. Dr. Lu has authored/coauthored more than 70 peerreviewed research papers and has led over a dozen patents and patent
applications.

Yang-Kook Sun received his M.S. degree and Ph.D. degree from the
Seoul National University, Korea. In 1996 he was a principal
researcher at Samsung SDI and contributed to the commercialization
of the lithium polymer battery. He has worked at the Hanyang
University in Korea as a professor since 2000. His research interests
are the synthesis of new electrode materials for lithium ion batteries,
supercapacitors, and next-generation batteries of Na ion batteries, Li
S batteries, and Liair batteries. In 2007 and 2011, he was awarded the
Energy Technology Division Research Award and Research Award in
Battery Division of the Electrochemical Society. He has been Senior
Visiting Scientist at the Argonne National Laboratory since 2007.

Li Li is an associate professor in the School of Chemical Engineering


and the Environment, Beijing Institute of Technology (BIT), China.
Her research interests focus on electrochemical energy storage and
conversion technology and the recycling technique and life-cycle
analysis for spent secondary batteries. As the principal investigator, Dr.
Li successfully hosted the National Key Program for Basic Research of
China, National High Tech 863 project, National Natural Science
Foundation, etc. Dr. Li earned her Ph.D. from BIT in 2004. She has
authored/coauthored over 60 research papers and led 18 patents and
patent applications in China.
Y

dx.doi.org/10.1021/cr400573b | Chem. Rev. XXXX, XXX, XXXXXX

Chemical Reviews

Review

also thankful for useful discussions with Dr. David Howell and
Mr. Tien Duong of the EERE Vehicle Technologies Oce.

Feng Wu was born in Beijing in 1951 and has been working at the
Beijing Institute of Technology (BIT) since graduating in 1982. Prof.
Wu is the director of the Professor Commission of the School of
Chemical Engineering and the Environment, the director of the Green
Energy Research Institute, and the vice president of the China Battery
Industry Association. His research interests are new electrode materials
and electrolytes for green secondary batteries. As the chief scientist, he
has hosted the National Key Program for Basic Research of China
(973 projects), Basic Research on New Rechargeable Batteries and
Related Materials, since 2002. He has published over 500 papers and
has been awarded 45 patents. In 2012, he was awarded the IBA
Research Award for outstanding research of battery material and
systems for electric vehicles.

REFERENCES
(1) Metz, B.; Davidson, O.; Coninck, H. C. d.; Loss, M.; Meyer., L. A.
IPCC Special Report on Carbon Dioxide Capture and Storage;
Cambridge University Press: Cambridge, U.K., New York, 2005.
(2) U.S. Energy Information Administration. Annual Energy Review
2010; Washington, DC, Oct 19, 2011.
(3) U.S. Census Bureau. Foreign Trade Statistics, 2011. http://www.
census.gov/foreign-trade/index.html (accessed May 8, 2013).
(4) Richter, B.; Goldston, D.; Crabtree, G.; Glicksman, L.; Goldstein,
D.; Greene, D.; Kammen, D.; Levine, M.; Lubell, M.; Savitz, M.;
Sperling, D. Energy Future: Think Eciency; American Physical Society:
College Park, MD, 2008.
(5) U.S. Environmental Protection Agency. Inventory of U.S.
Greenhouse Gas Emissions and Sinks 19902010; Washington, DC,
2012.
(6) Elgowainy, A.; Han, J.; Poch, L.; Wang, M.; Vyas, A.; Mahalik, M.;
Rousseau, A. Well-to-Wheels Analysis of Energy Use and Greenhouse Gas
Emissions of Plug-in Hybrid Electric Vehicles; Report No. ANL/ESD/092; Argonne National Laboratory: Argonne, IL, February 2009.
(7) Eberhard, M.; Tarpenning, M. The 21st Century Electric Car;
Tesla Motors: Palo Alto, CA, Oct 6, 2006.
(8) Yang, Z.; Zhang, J.; Kintner-Meyer, M. C. W.; Lu, X.; Choi, D.;
Lemmon, J. P.; Liu, J. Chem. Rev. 2011, 111, 3577.
(9) Bruce, P. G.; Freunberger, S. A.; Hardwick, L. J.; Tarascon, J.-M.
Nat. Mater. 2012, 11, 19.
(10) Abraham, K. M.; Jiang, Z. J. Electrochem. Soc. 1996, 143, 1.
(11) Bennion, D. N.; Littauer, E. L. J. Electrochem. Soc. 1976, 123,
1462.
(12) Littauer, E. L.; Tsai, K. C. J. Electrochem. Soc. 1976, 123, 771.
(13) Littauer, E. L.; Tsai, K. C. J. Electrochem. Soc. 1976, 123, 964.
(14) Dey, A. N. Thin Solid Films 1977, 43, 131.
(15) Littauer, E. L.; Tsai, K. C. J. Electrochem. Soc. 1977, 124, 850.
(16) Littauer, E. L.; Tsai, K. C.; Hollandsworth, R. P. J. Electrochem.
Soc. 1978, 125, 845.
(17) Blurton, K. F.; Sammells, A. F. J. Power Sources 1979, 4, 263.
(18) Gibian, M. J.; Sawyer, D. T.; Ungermann, T.;
Tangpoonpholvivat, R.; Morrison, M. M. J. Am. Chem. Soc. 1979,
101, 640.
(19) Peled, E. J. Electrochem. Soc. 1979, 126, 2047.
(20) Littauer, E. L.; Tsai, K. C. J. Electrochem. Soc. 1980, 127, 521.
(21) Momyer, W. R.; Littauer, E. L. Energy to the 21st Century,
Proceedings of the 15th Intersociety Energy Conversion Engineering
Conference, Seattle, WA, Aug 1822,1980; American Institute of
Aeronautics and Astronautics in cooperation with the IECEC Steering
Committee: New York, 1980; p 1480.
(22) Shao, Y.; Park, S.; Xiao, J.; Zhang, J.-G.; Wang, Y.; Liu, J. ACS
Catal. 2012, 2, 844.
(23) Black, R.; Adams, B.; Nazar, L. F. Adv. Energy Mater. 2012, 2,
801.
(24) Cao, R.; Lee, J.-S.; Liu, M.; Cho, J. Adv. Energy Mater. 2012, 2,
816.
(25) Lee, J.-S.; Kim, S. T.; Cao, R.; Choi, N.-S.; Liu, M.; Lee, K. T.;
Cho, J. Adv. Energy Mater. 2011, 1, 34.
(26) Park, M.; Sun, H.; Lee, H.; Lee, J.; Cho, J. Adv. Energy Mater.
2012, 2, 780.
(27) Gallant, B. M.; Kwabi, D. G.; Mitchell, R. R.; Zhou, J.;
Thompson, C. V.; Shao-Horn, Y. Energy Environ. Sci. 2013, 6, 2518.
(28) Cheng, F.; Liang, J.; Tao, Z.; Chen, J. Adv. Mater. 2011, 23,
1695.
(29) Ellis, B. L.; Lee, K. T.; Nazar, L. F. Chem. Mater. 2010, 22, 691.
(30) Cheng, F.; Chen, J. Chem. Soc. Rev. 2012, 41, 2172.
(31) Jeong, G.; Kim, Y.-U.; Kim, H.; Kim, Y.-J.; Sohn, H.-J. Energy
Environ. Sci. 2011, 4, 1986.
(32) Li, F.; Zhang, T.; Zhou, H. Energy Environ. Sci. 2013, 6, 1125.
(33) Beattie, S. D.; Manolescu, D. M.; Blair, S. L. J. Electrochem. Soc.
2009, 156, A44.

Khalil Amine is a Distinguished Fellow and Manager of the Advanced


Battery Technology programs at Argonne National Laboratory, where
he is responsible for directing the research and development of
advanced materials and battery systems for HEV, PHEV, EV, satellite,
military, and medical applications. Dr. Amine currently serves as a
member of the U.S. National Research Council on battery-related
technologies. Among his many awards, Dr. Amine is a 2003 recipient
of Scientic Americas Top Worldwide Research 50 Research Award, a
2009 recipient of the U.S. Federal Laboratory Award for Excellence in
Technology Transfer, and a four-time recipient of the R&D 100
Award, which is considered as the Oscar of technology and innovation.
In addition, he was recently awarded the ECS Battery Technology
Award and the International Battery Association Award. Dr. Amine
holds or has led over 130 patents and patent applications and has
over 280 publications. From 1998 to 2008, Dr. Amine was the most
cited scientist in the world in the eld of battery technology.

ACKNOWLEDGMENTS
This work was supported by the U.S. Department of Energy
(DOE) under Contract DE-AC0206CH11357, with the main
support provided by the Vehicle Technologies Oce, DOE
Oce of Energy Eciency and Renewable Energy (EERE). J.L.
was supported by a DOE EERE Postdoctoral Research Award
under the EERE Vehicles Technology Program administered by
the Oak Ridge Institute for Science and Education (ORISE) for
the DOE. This work was also partially supported by the
International S&T Cooperation Program of China (Grant
2010DFB63370) and the Chinese National 973 Program
(Grant 2009CB220100). This work was supported by the
Human Resources Development program (No.
20124010203310) of the Korea Institute of Energy Technology
Evaluation and Planning (KETEP) grant funded by the Korea
government Ministry of Trade, Industry and Energy and by a
National Research Foundation (NRF) of Korea grant funded
by the Korean government (MEST) (2009-0092780). We are
Z

dx.doi.org/10.1021/cr400573b | Chem. Rev. XXXX, XXX, XXXXXX

Chemical Reviews

Review

(71) Yang, Y.; Sun, Q.; Li, Y.-S.; Li, H.; Fu, Z.-W. J. Electrochem. Soc.
2011, 158, B1211.
(72) Read, J.; Motolo, K.; Ervin, M.; Behl, W.; Wolfenstine, J.;
Driedger, A.; Foster, D. J. Electrochem. Soc. 2003, 150, A1351.
(73) Aurbach, D.; Gofer, Y.; Langzam, J. J. Electrochem. Soc. 1989,
136, 3198.
(74) Aurbach, D.; Daroux, M.; Faguy, P.; Yeager, E. J. Electroanal.
Chem. 1991, 297, 225.
(75) Mizuno, F.; Nakanishi, S.; Kotani, Y.; Yokoishi, S.; Iba, H.
Electrochemistry (Tokyo, Jpn.) 2010, 78, 403.
(76) Xu, W.; Xu, K.; Viswanathan, V. V.; Towne, S. A.; Hardy, J. S.;
Xiao, J.; Hu, D.; Wang, D.; Zhang, J.-G. J. Power Sources 2011, 196,
9631.
(77) Xu, W.; Viswanathan, V. V.; Wang, D.; Towne, S. A.; Xiao, J.;
Nie, Z.; Hu, D.; Zhang, J.-G. J. Power Sources 2011, 196, 3894.
(78) McCloskey, B. D.; Bethune, D. S.; Shelby, R. M.; Girishkumar,
G.; Luntz, A. C. J. Phys. Chem. Lett. 2011, 2, 1161.
(79) Freunberger, S. A.; Chen, Y.; Peng, Z.; Griffin, J. M.; Hardwick,
L. J.; Barde, F.; Novak, P.; Bruce, P. G. J. Am. Chem. Soc. 2011, 133,
8040.
(80) Bryantsev, V. S.; Giordani, V.; Walker, W.; Blanco, M.; Zecevic,
S.; Sasaki, K.; Uddin, J.; Addison, D.; Chase, G. V. J. Phys. Chem. A
2011, 115, 12399.
(81) Bryantsev, V. S.; Blanco, M. J. Phys. Chem. Lett. 2011, 2, 379.
(82) Laino, T.; Curioni, A. Chem.Eur. J. 2012, 18, 3510.
(83) Freunberger, S. A.; Chen, Y.; Drewett, N. E.; Hardwick, L. J.;
Barde, F.; Bruce, P. G. Angew. Chem., Int. Ed. 2011, 50, 8609.
(84) Peng, Z.; Freunberger, S. A.; Hardwick, L. J.; Chen, Y.;
Giordani, V.; Barde, F.; Novak, P.; Graham, D.; Tarascon, J.-M.; Bruce,
P. G. Angew. Chem., Int. Ed. 2011, 50, 6351.
(85) Xu, D.; Wang, Z.-l.; Xu, J.-j.; Zhang, L.-l.; Zhang, X.-b. Chem.
Commun. 2012, 48, 6948.
(86) Mizuno, F.; Nakanishi, S.; Shirasawa, A.; Takechi, K.; Shiga, T.;
Nishikoori, H.; Iba, H. Electrochemistry (Tokyo, Jpn.) 2011, 79, 876.
(87) Lu, Y.-C.; Kwabi, D. G.; Yao, K. P. C.; Harding, J. R.; Zhou, J.;
Zuin, L.; Shao-Horn, Y. Energy Environ. Sci. 2011, 4, 2999.
(88) McCloskey, B. D.; Scheffler, R.; Speidel, A.; Bethune, D. S.;
Shelby, R. M.; Luntz, A. C. J. Am. Chem. Soc. 2011, 133, 18038.
(89) Xiao, J.; Hu, J.; Wang, D.; Hu, D.; Xu, W.; Graff, G. L.; Nie, Z.;
Liu, J.; Zhang, J.-G. J. Power Sources 2011, 196, 5674.
(90) Xu, W.; Hu, J.; Engelhard, M. H.; Towne, S. A.; Hardy, J. S.;
Xiao, J.; Feng, J.; Hu, M. Y.; Zhang, J.; Ding, F.; Gross, M. E.; Zhang,
J.-G. J. Power Sources 2012, 215, 240.
(91) Jung, H.-G.; Hassoun, J.; Park, J.-B.; Sun, Y.-K.; Scrosati, B. Nat.
Chem. 2012, 4, 579.
(92) Assary, R. S.; Curtiss, L. A.; Redfern, P. C.; Zhang, Z.; Amine, K.
J. Phys. Chem. C 2011, 115, 12216.
(93) Assary, R. S.; Lau, K. C.; Amine, K.; Sun, Y.-K.; Curtiss, L. A. J.
Phys. Chem. C 2013, 117, 8041.
(94) Bryantsev, V. S.; Faglioni, F. J. Phys. Chem. C 2012, 116, 7128.
(95) Bryantsev, V. S.; Uddin, J.; Giordani, V.; Walker, W.; Addison,
D.; Chase, G. V. J. Electrochem. Soc. 2013, 160, A160.
(96) McCloskey, B. D.; Bethune, D. S.; Shelby, R. M.; Mori, T.;
Scheffler, R.; Speidel, A.; Sherwood, M.; Luntz, A. C. J. Phys. Chem.
Lett. 2012, 3, 3043.
(97) Zhu, D.; Zhang, L.; Song, M.; Wang, X.; Mei, J.; Lau, L. M.;
Chen, Y. J. Solid State Electrochem. 2013, 1.
(98) Wang, H.; Xie, K. Electrochim. Acta 2012, 64, 29.
(99) Lim, H.-K.; Lim, H.-D.; Park, K.-Y.; Seo, D.-H.; Gwon, H.;
Hong, J.; Goddard, W. A.; Kim, H.; Kang, K. J. Am. Chem. Soc. 2013,
135, 9733.
(100) Ren, X.; Wu, Y. J. Am. Chem. Soc. 2013, 135, 2923.
(101) Mitchell, R. R.; Gallant, B. M.; Shao-Horn, Y.; Thompson, C.
V. J. Phys. Chem. Lett. 2013, 4, 1060.
(102) Gallant, B. M.; Mitchell, R. R.; Kwabi, D. G.; Zhou, J.; Zuin, L.;
Thompson, C. V.; Shao-Horn, Y. J. Phys. Chem. C 2012, 116, 20800.
(103) Cui, Y.; Wen, Z.; Liang, X.; Lu, Y.; Jin, J.; Wu, M.; Wu, X.
Energy Environ. Sci. 2012, 5, 7893.

(34) Christensen, J.; Albertus, P.; Sanchez-Carrera, R. S.; Lohmann,


T.; Kozinsky, B.; Liedtke, R.; Ahmed, J.; Kojic, A. J. Electrochem. Soc.
2012, 159, R1.
(35) Girishkumar, G.; McCloskey, B.; Luntz, A. C.; Swanson, S.;
Wilcke, W. J. Phys. Chem. Lett. 2010, 1, 2193.
(36) Capsoni, D.; Bini, M.; Ferrari, S.; Quartarone, E.; Mustarelli, P.
J. Power Sources 2012, 220, 253.
(37) Kraytsberg, A.; Yair, E.-E. J. Power Sources 2011, 196, 886.
(38) Padbury, R.; Zhang, X. J. Power Sources 2011, 196, 4436.
(39) Song, M.-K.; Park, S.; Alamgir, F. M.; Cho, J.; Liu, M. Mater. Sci.
Eng., R 2011, 72, 203.
(40) Bruce, P. G.; Hardwick, L. J.; Abraham, K. M. MRS Bull. 2011,
36, 506.
(41) Kowalczk, I.; Read, J.; Salomon, M. Pure Appl. Chem. 2007, 79,
851.
(42) Shao, Y.; Ding, F.; Xiao, J.; Zhang, J.; Xu, W.; Park, S.; Zhang, J.G.; Wang, Y.; Liu, J. Adv. Funct. Mater. 2013, 23, 987.
(43) Sun, Y. Nano Energy 2013, 2, 801.
(44) Lei, Y.; Lu, J.; Luo, X.; Wu, T.; Du, P.; Zhang, X.; Ren, Y.;
Miller, J. T.; Sun, Y.-K.; Elam, J. W. Nano Lett. 2013, 13, 4182.
(45) Qin, Y.; Lu, J.; Du, P.; Chen, Z.; Ren, Y.; Wu, T.; Miller, J. T.;
Wen, J.; Miller, D. J.; Zhang, Z.; Amine, K. Energy Environ. Sci. 2013, 6,
519.
(46) Adams, B. D.; Radtke, C.; Black, R.; Trudeau, M. L.; Zaghib, K.;
Nazar, L. F. Energy Environ. Sci. 2013, 6, 1772.
(47) Lu, J.; Qin, Y.; Du, P.; Luo, X.; Wu, T.; Ren, Y.; Wen, J.; Miller,
D. J.; Miller, J. T.; Amine, K. RSC Adv. 2013, 3, 8276.
(48) Zhang, J.-G.; Wang, D.; Xu, W.; Xiao, J.; Williford, R. E. J. Power
Sources 2010, 195, 4332.
(49) Wang, Y.; Zhou, H. Energy Environ. Sci. 2011, 4, 1704.
(50) Zheng, J. P.; Liang, R. Y.; Hendrickson, M.; Plichta, E. J. J.
Electrochem. Soc. 2008, 155, A432.
(51) Ogasawara, T.; Debart, A.; Holzapfel, M.; Novak, P.; Bruce, P.
G. J. Am. Chem. Soc. 2006, 128, 1390.
(52) He, P.; Wang, Y.; Zhou, H. Electrochem. Commun. 2010, 12,
1686.
(53) Kumar, B.; Kumar, J. J. Electrochem. Soc. 2010, 157, A611.
(54) Wang, Y.; Zhou, H. J. Power Sources 2010, 195, 358.
(55) Zhang, Z. C.; 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. J. Phys. Chem. C 2011, 115, 25535.
(56) Debart, A.; Paterson, A. J.; Bao, J.; Bruce, P. G. Angew. Chem.,
Int. Ed. 2008, 47, 4521.
(57) Assary, R. S.; Lu, J.; Du, P.; Luo, X.; Zhang, X.; Ren, Y.; Curtiss,
L. A.; Amine, K. ChemSusChem 2013, 6, 51.
(58) Xu, W.; Xiao, J.; Zhang, J.; Wang, D.; Zhang, J.-G. J. Electrochem.
Soc. 2009, 156, A773.
(59) Peng, Z.; Freunberger, S. A.; Chen, Y.; Bruce, P. G. Science 2012,
337, 563.
(60) Laoire, C. O.; Mukerjee, S.; Plichta, E. J.; Hendrickson, M. A.;
Abraham, K. M. J. Electrochem. Soc. 2011, 158, A302.
(61) Mitchell, R. R.; Gallant, B. M.; Thompson, C. V.; Shao-Horn, Y.
Energy Environ. Sci. 2011, 4, 2952.
(62) Xiao, J.; Mei, D.; Li, X.; Xu, W.; Wang, D.; Graff, G. L.; Bennett,
W. D.; Nie, Z.; Saraf, L. V.; Aksay, I. A.; Liu, J.; Zhang, J.-G. Nano Lett.
2011, 11, 5071.
(63) Lu, Y.-C.; Xu, Z.; Gasteiger, H. A.; Chen, S.; Hamad-Schifferli,
K.; Shao-Horn, Y. J. Am. Chem. Soc. 2010, 132, 12170.
(64) Laoire, C. O.; Mukerjee, S.; Abraham, K. M.; Plichta, E. J.;
Hendrickson, M. A. J. Phys. Chem. C 2009, 113, 20127.
(65) Read, J. J. Electrochem. Soc. 2002, 149, A1190.
(66) Read, J.; Mutolo, K.; Ervin, M.; Behl, W.; Wolfenstine, J.;
Driedger, A.; Foster, D. J. Electrochem. Soc. 2003, 150, A1351.
(67) Read, J. J. Electrochem. Soc. 2006, 153, A96.
(68) Zhang, S. S.; Foster, D.; Read, J. J. Power Sources 2010, 195,
1235.
(69) Xu, K. Chem. Rev. 2004, 104, 4303.
(70) Thapa, A. K.; Ishihara, T. J. Power Sources 2011, 196, 7016.
AA

dx.doi.org/10.1021/cr400573b | Chem. Rev. XXXX, XXX, XXXXXX

Chemical Reviews

Review

(104) Harding, J. R.; Lu, Y.-C.; Tsukada, Y.; Shao-Horn, Y. Phys.


Chem. Chem. Phys. 2012, 14, 10540.
(105) Lu, Y.-C.; Gasteiger, H. A.; Shao-Horn, Y. J. Am. Chem. Soc.
2011, 133, 19048.
(106) Lu, Y.-C.; Shao-Horn, Y. J. Phys. Chem. Lett. 2013, 4, 93.
(107) Wu, G.; Mack, N. H.; Gao, W.; Ma, S.; Zhong, R.; Han, J.;
Baldwin, J. K.; Zelenay, P. ACS Nano 2012, 6, 9764.
(108) Shui, J.-L.; Okasinski, J. S.; Kenesei, P.; Dobbs, H. A.; Zhao, D.;
Almer, J. D.; Liu, D.-J. Nat. Commun. 2013, 4, 3255.
(109) Lei, Y.; Lu, J.; Luo, X.; Wu, T.; Du, P.; Zhang, X.; Ren, Y.;
Wen, J.; Miller, D. J.; Miller, J. T.; Sun, Y.-K.; Elam, J. W.; Amine, K.
Nano Lett. 2013, 13, 4182.
(110) Zhang, L.; Zhang, S.; Zhang, K.; Xu, G.; He, X.; Dong, S.; Liu,
Z.; Huang, C.; Gu, L.; Cui, G. Chem. Commun. 2013, 49, 3540.
(111) Xu, J.-J.; Xu, D.; Wang, Z.-L.; Wang, H.-G.; Zhang, L.-L.;
Zhang, X.-B. Angew. Chem., Int. Ed. 2013, 52, 3887.
(112) Trahan, M. J.; Jia, Q.; Mukerjee, S.; Plichta, E. J.; Hendrickson,
M. A.; Abraham, K. M. J. Electrochem. Soc. 2013, 160, A1577.
(113) Shui, J.-L.; Wang, H.-H.; Liu, D.-J. Electrochem. Commun. 2013,
34, 45.
(114) Shui, J.; Okasinski, J.; Zhao, D.; ALmer, J.; Liu, D.-J. ESC
Trans. 2013, 50, 37.
(115) Park, H. W.; Lee, D. U.; Nazar, L. F.; Chen, Z. J. Electrochem.
Soc. 2013, 160, A344.
(116) Nasybulin, E.; Xu, W.; Engelhard, M. H.; Li, X. S.; Gu, M.; Hu,
D.; Zhang, J.-G. Electrochem. Commun. 2013, 29, 63.
(117) Meini, S.; Tsiouvaras, N.; Schwenke, K. U.; Piana, M.; Beyer,
H.; Lange, L.; Gasteiger, H. A. Phys. Chem. Chem. Phys. 2013, 15,
11478.
(118) Lim, H.-D.; Park, K.-Y.; Song, H.; Jang, E. Y.; Gwon, H.; Kim,
J.; Kim, Y. H.; Lima, M. D.; Robles, R. O.; Lepro, X.; Baughman, R. H.;
Kang, K. Adv. Mater. 2013, 25, 1348.
(119) Li, F.; Ohnishi, R.; Yamada, Y.; Kubota, J.; Domen, K.;
Yamada, A.; Zhou, H. Chem. Commun. 2013, 49, 1175.
(120) Jung, H.-G.; Jeong, Y. S.; Park, J.-B.; Sun, Y.-K.; Scrosati, B.;
Lee, Y. J. ACS Nano 2013, 7, 3532.
(121) Huang, B.-W.; Liao, X.-Z.; Wang, H.; Wang, C.-N.; He, Y.-S.;
Ma, Z.-F. J. Electrochem. Soc. 2013, 160, A1112.
(122) Elia, G. A.; Park, J.-B.; Scrosati, B.; Sun, Y.-K.; Hassoun, J.
Electrochem. Commun. 2013, 34, 250.
(123) Barile, C. J.; Gewirth, A. A. J. Electrochem. Soc. 2013, 160, A549.
(124) Wu, J.; Park, H. W.; Yu, A.; Higgins, D.; Chen, Z. J. Phys.
Chem. C 2012, 116, 9427.
(125) Truong, T. T.; Liu, Y.; Ren, Y.; Trahey, L.; Sun, Y. ACS Nano
2012, 6, 8067.
(126) Fu, Z.; Lin, X.; Huang, T.; Yu, A. J. Solid State Electrochem.
2012, 16, 1447.
(127) Shui, J.-L.; Karan, N. K.; Balasubramanian, M.; Li, S.-Y.; Liu,
D.-J. J. Am. Chem. Soc. 2012, 134, 16654.
(128) Bryantsev, V. S. Theor. Chem. Acc. 2012, 131, 1.
(129) Oloniyo, O.; Kumar, S.; Scott, K. J. Electron. Mater. 2012, 41,
921.
(130) Adams, J.; Karulkar, M. J. Power Sources 2012, 199, 247.
(131) Lim, H.-D.; Park, K.-Y.; Gwon, H.; Hong, J.; Kim, H.; Kang, K.
Chem. Commun. 2012, 48, 8374.
(132) Li, Y.; Wang, J.; Li, X.; Geng, D.; Banis, M. N.; Li, R.; Sun, X.
Electrochem. Commun. 2012, 18, 12.
(133) Lee, S. W.; Gallant, B. M.; Lee, Y.; Yoshida, N.; Kim, D. Y.;
Yamada, Y.; Noda, S.; Yamada, A.; Shao-Horn, Y. Energy Environ. Sci.
2012, 5, 5437.
(134) Hassoun, J.; Jung, H.-G.; Lee, D.-J.; Park, J.-B.; Amine, K.; Sun,
Y.-K.; Scrosati, B. Nano Lett. 2012, 12, 5775.
(135) Black, R.; Oh, S. H.; Lee, J.-H.; Yim, T.; Adams, B.; Nazar, L. F.
J. Am. Chem. Soc. 2012, 134, 2902.
(136) Jung, H.-G.; Kim, H.-S.; Park, J.-B.; Oh, I.-H.; Hassoun, J.;
Yoon, C. S.; Scrosati, B.; Sun, Y.-K. Nano Lett. 2012, 12, 4333.
(137) Zhang, Z. C.; Lyons, L. J.; Jin, J. J.; Amine, K.; West, R. Chem.
Mater. 2005, 17, 5646.

(138) Ma, Y.; Li, N.; Li, D.; Zhang, M.; Huang, X. J. Power Sources
2011, 196, 2346.
(139) Chen, Z.; Wang, H. H.; Vissers, D. R.; Zhang, L.; West, R.;
Lyons, L. J.; Amine, K. J. Phys. Chem. C 2008, 112, 2210.
(140) Oh, B.; Vissers, D.; Zhang, Z.; West, R.; Tsukamoto, H.;
Amine, K. J. Power Sources 2003, 119, 442.
(141) Laoire, C. O.; Mukerjee, S.; Abraham, K. M.; Plichta, E. J.;
Hendrickson, M. A. J. Phys. Chem. C 2010, 114, 9178.
(142) Lopez, N.; Graham, D. J.; McGuire, R.; Alliger, G. E.; ShaoHorn, Y.; Cummins, C. C.; Nocera, D. G. Science 2012, 335, 450.
(143) Trahan, M. J.; Mukerjee, S.; Plichta, E. J.; Hendrickson, M. A.;
Abraham, K. M. J. Electrochem. Soc. 2013, 160, A259.
(144) Sun, B.; Zhang, J.; Munroe, P.; Ahn, H.-J.; Wang, G.
Electrochem. Commun. 2013, 31, 88.
(145) Ottakam Thotiyl, M. M.; Freunberger, S. A.; Peng, Z.; Bruce,
P. G. J. Am. Chem. Soc. 2012, 135, 494.
(146) Mozhzhukhina, N.; Mendez De Leo, L. P.; Calvo, E. J. J. Phys.
Chem. C 2013, 117, 18375.
(147) Chen, Y.; Freunberger, S. A.; Peng, Z.; Fontaine, O.; Bruce, P.
G. Nat. Chem. 2013, 5, 489.
(148) McCloskey, B. D.; Valery, A.; Luntz, A. C.; Gowda, S. R.;
Wallraff, G. M.; Garcia, J. M.; Mori, T.; Krupp, L. E. J. Phys. Chem. Lett.
2013, 2989.
(149) Chen, Y.; Freunberger, S. A.; Peng, Z.; Barde, F.; Bruce, P. G. J.
Am. Chem. Soc. 2012, 134, 7952.
(150) Walker, W.; Giordani, V.; Uddin, J.; Bryantsev, V. S.; Chase, G.
V.; Addison, D. J. Am. Chem. Soc. 2013, 135, 2076.
(151) Giordani, V.; Walker, W.; Bryantsev, V. S.; Uddin, J.; Chase, G.
V.; Addison, D. J. Electrochem. Soc. 2013, 160, A1544.
(152) Crowther, O.; Meyer, B.; Salomon, M. Electrochem. Solid-State
Lett. 2011, 14, A113.
(153) Allen, C. J.; Mukerjee, S.; Plichta, E. J.; Hendrickson, M. A.;
Abraham, K. M. J. Phys. Chem. Lett. 2011, 2, 2420.
(154) De Giorgio, F.; Soavi, F.; Mastragostino, M. Electrochem.
Commun. 2011, 13, 1090.
(155) Zhang, Z.; Zhang, L.; Schlueter, J. A.; Redfern, P. C.; Curtiss,
L.; Amine, K. J. Power Sources 2010, 195, 4957.
(156) Zhang, L.; Zhang, Z.; Wu, H.; Amine, K. Energy Environ. Sci.
2011, 4, 2858.
(157) Zhang, L.; Zhang, Z.; Redfern, P. C.; Curtiss, L. A.; Amine, K.
Energy Environ. Sci. 2012, 5, 8204.
(158) Qin, Y.; Chen, Z.; Liu, J.; Amine, K. Electrochem. Solid-State
Lett. 2010, 13, A11.
(159) Chen, Z.; Ren, Y.; Jansen, A. N.; Lin, C.-k.; Weng, W.; Amine,
K. Nat. Commun. 2013, 4, No. 1513.
(160) Chen, Z.; Qin, Y.; Amine, K. Electrochim. Acta 2009, 54, 5605.
(161) Chen, Z.; Liu, J.; Jansen, A. N.; GirishKumar, G.; Casteel, B.;
Amine, K. Electrochem. Solid-State Lett. 2010, 13, A39.
(162) Kichambare, P.; Rodrigues, S.; Kumar, J. ACS Appl. Mater.
Interfaces 2012, 4, 49.
(163) Kumar, A.; Ciucci, F.; Morozovska, A. N.; Kalinin, S. V.; Jesse,
S. Nat. Chem. 2011, 3, 707.
(164) Zhang, D.; Li, R.; Huang, T.; Yu, A. J. Power Sources 2010, 195,
1202.
(165) Hassoun, J.; Croce, F.; Armand, M.; Scrosati, B. Angew. Chem.,
Int. Ed. 2011, 50, 2999.
(166) Kamaya, N.; Homma, K.; Yamakawa, Y.; Hirayama, M.; Kanno,
R.; Yonemura, M.; Kamiyama, T.; Kato, Y.; Hama, S.; Kawamoto, K.;
Mitsui, A. Nat. Mater. 2011, 10, 682.
(167) Cecchetto, L.; Salomon, M.; Scrosati, B.; Croce, F. J. Power
Sources 2012, 213, 233.
(168) Guyomard, D.; Tarascon, J. M. J. Electrochem. Soc. 1993, 140,
3071.
(169) Tarascon, J. M.; Guyomard, D. Solid State Ionics 1994, 69, 293.
(170) Oswald, S.; Mikhailova, D.; Scheiba, F.; Reichel, P.; Fiedler, A.;
Ehrenberg, H. Anal. Bioanal. Chem. 2011, 400, 691.
(171) Veith, G. M.; Dudney, N. J.; Howe, J.; Nanda, J. J. Phys. Chem.
C 2011, 115, 14325.
AB

dx.doi.org/10.1021/cr400573b | Chem. Rev. XXXX, XXX, XXXXXX

Chemical Reviews

Review

(172) Du, P.; Lu, J.; Lau, K. C.; Luo, X.; Bareno, J.; Zhang, X.; Ren,
Y.; Zhang, Z.; Curtiss, L. A.; Sun, Y.-K.; Amine, K. Phys. Chem. Chem.
Phys. 2013, 15, 5572.
(173) Pearson, R. G. J. Am. Chem. Soc. 1963, 85, 3533.
(174) Yang, J.; Zhai, D.; Wang, H.-H.; Lau, K. C.; Schlueter, J. A.;
Du, P.; Myers, D. J.; Sun, Y.-K.; Curtiss, L. A.; Amine, K. Phys. Chem.
Chem. Phys. 2013, 15, 3764.
(175) Black, R.; Lee, J.-H.; Adams, B.; Mims, C. A.; Nazar, L. F.
Angew. Chem., Int. Ed. 2013, 52, 392.
(176) Nelson, R.; Weatherspoon, M. H.; Gomez, J.; Kalu, E. E.;
Zheng, J. P. Electrochem. Commun. 2013, 34, 77.
(177) Zhang, G. Q.; Hendrickson, M.; Plichta, E. J.; Au, M.; Zheng, J.
P. J. Electrochem. Soc. 2012, 159, A310.
(178) Lee, B. G.; Nam, S.-C.; Choi, J. Curr. Appl. Phys. 2012, 12,
1580.
(179) Zhang, Y.; Zhang, H.; Li, J.; Wang, M.; Nie, H.; Zhang, F. J.
Power Sources 2013, 240, 390.
(180) Kim, K. S.; Park, Y. J. Nanoscale Res. Lett. 2012, 7, 47.
(181) Kalubarme, R. S.; Park, C.-J. Chem. Lett. 2012, 41, 612.
(182) Adams, J.; Karulkar, M.; Anandan, V. J. Power Sources 2013,
239, 132.
(183) Meini, S.; Piana, M.; Tsiouvaras, N.; Garsuch, A.; Gasteiger, H.
A. Electrochem. Solid-State Lett. 2012, 15, A45.
(184) Crowther, O.; Salomon, M. Membranes 2012, 2, 216.
(185) Wang, H.; Liao, X.; Jiang, Q.; Yang, X.; He, Y.; Ma, Z. Chin. Sci.
Bull. 2012, 57, 1959.
(186) Selvaraj, C.; Munichandraiah, N.; Scanlon, L. G. J. Porphyrins
Phthalocyanines 2012, 16, 255.
(187) Zheng, D.; Lee, H.-S.; Yang, X.-Q.; Qu, D. Electrochem.
Commun. 2013, 28, 17.
(188) Garsuch, A.; Badine, D. M.; Leitner, K.; Gasparotto, L. H. S.;
Borisenko, N.; Endres, F.; Vracar, M.; Janek, J.; Oesten, R. Z. Phys.
Chem. 2012, 226, 107.
(189) Crowther, O.; Keeny, D.; Moureau, D. M.; Meyer, B.;
Salomon, M.; Hendrickson, M. J. Power Sources 2012, 202, 347.
(190) Zhang, L.; Zhang, X.; Wang, Z.; Xu, J.; Xu, D.; Wang, L. Chem.
Commun. 2012, 48, 7598.
(191) Tsiouvaras, N.; Meini, S.; Buchberger, I.; Gasteiger, H. A. J.
Electrochem. Soc. 2013, 160, A471.
(192) Yuasa, M.; Matsuyoshi, T.; Kida, T.; Shimanoe, K. J. Power
Sources 2013, 242, 216.
(193) Younesi, R.; Urbonaite, S.; Edstrom, K.; Hahlin, M. J. Phys.
Chem. C 2012, 116, 20673.
(194) Wang, H.; Yang, Y.; Liang, Y.; Zheng, G.; Li, Y.; Cui, Y.; Dai,
H. Energy Environ. Sci. 2012, 5, 7931.
(195) Sun, B.; Wang, B.; Su, D.; Xiao, L.; Ahn, H.; Wang, G. Carbon
2012, 50, 727.
(196) Sun, B.; Liu, H.; Munroe, P.; Ahn, H.; Wang, G. Nano Res.
2012, 5, 460.
(197) Shitta-Bey, G. O.; Mirzaeian, M.; Halla, P. J. J. Electrochem. Soc.
2012, 159, A315.
(198) Mirzaeian, M.; Hall, P. J.; Sillars, F. B.; Fletcher, I.; Goldin, M.
M.; Shitta-bey, G. O.; Jirandehi, H. F. J. Electrochem. Soc. 2013, 160,
A25.
(199) Herranz, J.; Garsuch, A.; Gasteiger, H. A. J. Phys. Chem. C
2012, 116, 19084.
(200) Yoon, T. H.; Park, Y. J. Nanoscale Res. Lett. 2012, 7, 1.
(201) Gao, Y.; Wang, C.; Pu, W.; Liu, Z.; Deng, C.; Zhang, P.; Mao,
Z. Int. J. Hydrogen Energy 2012, 37, 12725.
(202) 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, 715.
(203) Li, L.; Zhao, X.; Manthiram, A. Electrochem. Commun. 2012,
14, 78.
(204) Ke, F.-S.; Solomon, B. C.; Ma, S.-G.; Zhou, X.-D. Electrochim.
Acta 2012, 85, 444.
(205) Huff, L. A.; Rapp, J. L.; Zhu, L.; Gewirth, A. A. J. Power Sources
2013, 235, 87.

(206) Leskes, M.; Drewett, N. E.; Hardwick, L. J.; Bruce, P. G.;


Goward, G. R.; Grey, C. P. Angew. Chem., Int. Ed. 2012, 51, 8560.
(207) Zhamu, A.; Chen, G.; Liu, C.; Neff, D.; Fang, Q.; Yu, Z.;
Xiong, W.; Wang, Y.; Wang, X.; Jang, B. Z. Energy Environ. Sci. 2012, 5,
5701.
(208) Yin, J.; Fang, B.; Luo, J.; Wanjala, B.; Mott, D.; Loukrakpam,
R.; Ng, M. S.; Li, Z.; Hong, J.; Whittingham, M. S.; Zhong, C.-J.
Nanotechnology 2012, 23, 305404.
(209) Yang, Y.; Shi, M.; Zhou, Q.-F.; Li, Y.-S.; Fu, Z.-W. Electrochem.
Commun. 2012, 20, 11.
(210) Wang, L.; Ara, M.; Wadumesthrige, K.; Salley, S.; Ng, K. Y. S. J.
Power Sources 2013, 234, 8.
(211) He, H.; Niu, W.; Asl, N. M.; Salim, J.; Chen, R.; Kim, Y.
Electrochim. Acta 2012, 67, 87.
(212) Cao, Y.; Wei, Z.; He, J.; Zang, J.; Zhang, Q.; Zheng, M.; Dong,
Q. Energy Environ. Sci. 2012, 5, 9765.
(213) Yoo, E.; Nakamura, J.; Zhou, H. Energy Environ. Sci. 2012, 5,
6928.
(214) Thapa, A. K.; Shin, T. H.; Ida, S.; Sumanasekera, G. U.;
Sunkara, M. K.; Ishihara, T. J. Power Sources 2012, 220, 211.
(215) Ida, S.; Thapa, A. K.; Hidaka, Y.; Okamoto, Y.; Matsuka, M.;
Hagiwara, H.; Ishihara, T. J. Power Sources 2012, 203, 159.
(216) Trahey, L.; Karan, N. K.; Chan, M. K. Y.; Lu, J.; Ren, Y.;
Greeley, J.; Balasubramanian, M.; Burrell, A. K.; Curtiss, L. A.;
Thackeray, M. M. Adv. Energy Mater. 2013, 3, 75.
(217) Sun, B.; Munroe, P.; Wang, G. Sci. Rep. 2013, 3, 2247.
(218) Lau, K. C.; Curtiss, L. A.; Greeley, J. J. Phys. Chem. C 2011,
115, 23625.
(219) Chen, J.; Hummelshoj, J. S.; Thygesen, K. S.; Myrdal, J. S. G.;
Norskov, J. K.; Vegge, T. Catal. Today 2011, 165, 2.
(220) Garcia-Lastra, J. M.; Bass, J. D.; Thygesen, K. S. J. Chem. Phys.
2011, 135, 121101.
(221) Ong, S. P.; Mo, Y.; Ceder, G. Phys. Rev. B 2012, 85,
081105(R).
(222) Hummelshoj, J. S.; Blomqvist, J.; Datta, S.; Vegge, T.;
Rossmeisl, J.; Thygesen, K. S.; Luntz, A. C.; Jacobsen, K. W.; Norskov,
J. K. J. Chem. Phy. 2010, 132, 071101.
(223) Lau, K. C.; Assary, R. S.; Redfern, P.; Greeley, J.; Curtiss, L. A.
J. Phys. Chem. C 2012, 116, 23890.
(224) Radin, M. D.; Rodriguez, J. F.; Tian, F.; Siegel, D. J. J. Am.
Chem. Soc. 2012, 134, 1093.
(225) Lu, J.; Jung, H.-J.; Lau, K. C.; Zhang, Z.; Schlueter, J. A.; Du,
P.; Assary, R. S.; Greeley, J.; Ferguson, G. A.; Wang, H.-H.; Hassoun,
J.; Iddir, H.; Zhou, J.; Zuin, L.; Hu, Y.; Sun, Y.-K.; Scrosati, B.; Curtiss,
L. A.; Amine, K. ChemSusChem 2013, 6, 1196.
(226) Xiao, J.; Wang, D.; Xu, W.; Wang, D.; Williford, R. E.; Liu, J.;
Zhang, J.-G. J. Electrochem. Soc. 2010, 157, A487.
(227) Yang, X.-h.; He, P.; Xia, Y.-y. Electrochem. Commun. 2009, 11,
1127.
(228) Kuboki, T.; Okuyama, T.; Ohsaki, T.; Takami, N. J. Power
Sources 2005, 146, 766.
(229) Cheng, H.; Scott, K. Appl. Catal., B 2011, 108, 140.
(230) Debart, A.; Bao, J.; Armstrong, G.; Bruce, P. G. J. Power Sources
2007, 174, 1177.
(231) Jin, L.; Xu, L.; Morein, C.; Chen, C.-h.; Lai, M.; Dharmarathna,
S.; Dobley, A.; Suib, S. L. Adv. Funct. Mater. 2010, 20, 3373.
(232) Lee, S.; Zhu, S.; Milleville, C. C.; Lee, C.-Y.; Chen, P.;
Takeuchi, K. J.; Takeuchi, E. S.; Marschilok, A. C. Electrochem. SolidState Lett. 2010, 13, A162.
(233) Lu, Y.-C.; Gasteiger, H. A.; Crumlin, E.; McGuire, R., Jr.; ShaoHorn, Y. J. Electrochem. Soc. 2010, 157, A1016.
(234) Lu, Y.-C.; Gasteiger, H. A.; Parent, M. C.; Chiloyan, V.; ShaoHorn, Y. Electrochem. Solid-State Lett. 2010, 13, A69.
(235) Mirzaeian, M.; Hall, P. J. J. Power Sources 2010, 195, 6817.
(236) Thapa, A. K.; Saimen, K.; Ishihara, T. Electrochem. Solid-State
Lett. 2010, 13, A165.
(237) Wang, D.; Xiao, J.; Xu, W.; Zhang, J.-G. J. Electrochem. Soc.
2010, 157, A760.
(238) Wang, Y.; Zhou, H. Chem. Commun. 2010, 46, 6305.
AC

dx.doi.org/10.1021/cr400573b | Chem. Rev. XXXX, XXX, XXXXXX

Chemical Reviews

Review

(239) Xiao, J.; Xu, W.; Wang, D.; Zhang, J.-G. J. Electrochem. Soc.
2010, 157, A294.
(240) Yang, X.-h.; Xia, Y.-y. J. Solid State Electrochem. 2010, 14, 109.
(241) Zhang, D.; Fu, Z.; Wei, Z.; Huang, T.; Yu, A. J. Electrochem.
Soc. 2010, 157, A362.
(242) Zhang, G. Q.; Zheng, J. P.; Liang, R.; Zhang, C.; Wang, B.;
Hendrickson, M.; Plichta, E. J. J. Electrochem. Soc. 2010, 157, A953.
(243) Oh, S. H.; Black, R.; Pomerantseva, E.; Lee, J.-H.; Nazar, L. F.
Nat. Chem. 2012, 4, 1004.
(244) Ryu, W.-H.; Yoon, T.-H.; Song, S. H.; Jeon, S.; Park, Y.-J.; Kim,
I.-D. Nano Lett. 2013, 13, 4190.
(245) Trahey, L.; Johnson, C. S.; Vaughey, J. T.; Kang, S. H.;
Hardwick, L. J.; Freunberger, S. A.; Bruce, P. G.; Thackeray, M. M.
Electrochem. Solid-State Lett. 2011, 14, A64.
(246) 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.
(247) McCloskey, B. D.; Speidel, A.; Scheffler, R.; Miller, D. C.;
Viswanathan, V.; Hummelshoj, J. S.; Norskov, J. K.; Luntz, A. C. J.
Phys. Chem. Lett. 2012, 3, 997.
(248) Lu, J.; Lei, Y.; Lau, K. C.; Luo, X.; Du, P.; Wen, J.; Assary, R. S.;
Das, U.; Miller, D. J.; Elam, J. W.; Albishri, H. M.; El-Hady, D. A.; Sun,
Y.-K.; Curtiss, L. A.; Amine, K. Nat. Commun. 2013, 4, No. 2383.
(249) Aurbach, D.; Zinigrad, E.; Cohen, E.; Teller, H. Solid State
Ionics 2002, 148, 405.
(250) Yamaki, J.; Tobishima, S.; Hayashi, K.; Saito, K.; Nemoto, Y.;
Arakawa, M. J. Power Sources 1998, 74, 219.
(251) Zhang, T.; Imanishi, N.; Hasegawa, S.; Hirano, A.; Xie, J.;
Takeda, Y.; Yamamoto, O.; Sammes, N. J. Electrochem. Soc. 2008, 155,
A965.
(252) Kumar, B.; Kumar, J.; Leese, R.; Fellner, J. P.; Rodrigues, S. J.;
Abraham, K. M. J. Electroanal. Chem. 2010, 157, A50.
(253) Armand, M.; Tarascon, J.-M. Nature 2008, 451, 652.
(254) Truong, T. T.; Qin, Y.; Ren, Y.; Chen, Z.; Chan, M. K.;
Greeley, J. P.; Amine, K.; Sun, Y. Adv. Mater. 2011, 23, 4947.
(255) Wang, Y.; He, P.; Zhou, H. Energy Environ. Sci. 2011, 4, 4994.
(256) Zhang, J. G.; Wnag, D.; Xu, W.; Xiao, J.; Williford, R. E. J.
Power Sources 2010, 195, 4332.
(257) Gierszewski, P. J.; Finn, P. A.; Kirk, D. W. Fusion Eng. Des.
1990, 13, 59.
(258) Visco, S.; Jonghe, L. D.; Nimon, Y.; Petrov, A.; Pridatko, K.
Aqueous Lithium/Air Battery Cells. WO Patent Application WO/
2010/005, 2010.
(259) Welna, D. T.; Stone, D. A.; Allcock, H. R. Chem. Mater. 2006,
18, 4486.
(260) Visco, S.; Katz, B. D.; Nimon, Y.; Jonghe, L. D. U.S. Patent
7,491,458, Feb 17, 2009; see also other patents cited therein.
(261) Visco, S.; Katz, B. D.; Nimon, Y.; Jonghe, L. D. U.S. Patent
7,282,295, Oct 16, 2007.
(262) Fu, J. U. S. Patent 6,485,622, Nov 26, 2002.
(263) Fu, J. Solid State Ionics 1997, 96, 195.
(264) Fu, J. Solid State Ion. 1997, 104, 191.
(265) Crowthe, O.; Salomon, M. In Lithium Batteries; Scrosati, B.,
Abraham, K. M., Schalkwijk, W., Hassoum, J., Eds.; John Wiley & Sons
Inc.: Hoboken, NJ, 2012; Chapter 9.
(266) Zhang, T.; Imanishi, N.; Shimonishi, Y.; Hirano, A.; Takeda,
Y.; Yamamoto, O.; Sammes, N. Chem. Commun. 2010, 46, 1661.
(267) Zhang, T.; Imanishi, N.; Hasegawa, S.; Hirano, A.; Xie, J.;
Takeda, Y.; Yamamoto, O.; Sammes, N. Electrochem. Solid-State Lett.
2009, 12, A132.
(268) Hasegawa, S.; Imanishi, N.; Zhang, T.; Xie, J.; Hirano, A.;
Takeda, Y.; Yamamoto, O. J. Power Sources 2009, 189, 371.
(269) Yoo, E.; Zhou, H. ACS Nano 2011, 5, 3020.
(270) He, P.; Wang, Y.; Zhou, H. Chem. Commun. 2011, 47, 10701.
(271) Wang, L.; Zhao, X.; Lu, Y.; Xu, M.; Zhang, D.; Ruoff, R. S.;
Stevenson, K. J.; Goodenough, J. B. J. Electrochem. Soc. 2011, 158,
A1379.
(272) Suntivich, J.; Gasteiger, H. A.; Yabuuchi, N.; Nakanishi, H.;
Goodenough, J. B.; Shao-Horn, Y. Nat. Chem. 2011, 3, 546.

(273) Thotiyl, M. M. O.; Freunberger, S. A.; Peng, Z.; Chen, Y.; Liu,
Z.; Bruce, P. G. Nat. Mater. 2013, 12, 1050.
(274) Sun, Y. G.; Choi, W. M.; Jiang, H. Q.; Huang, Y. G. Y.; Rogers,
J. A. Nature Nanotechnol. 2006, 1, 201.
(275) Baddour-Hadjean, R.; Pereira-Ramos, J.-P. Chem. Rev. 2009,
110, 1278.
(276) Ryan, K. R.; Trahey, L.; Okasinski, J. S.; Burrell, A. K.; Ingram,
B. J. J. Mater. Chem. A 2013, 1, 6915.
(277) Karan, N. K.; Balasubramanian, M.; Fister, T. T.; Burrell, A. K.;
Du, P. J. Phys. Chem. C 2012, 116, 18132.
(278) Chan, M. K. Y.; Shirley, E. L.; Karan, N. K.; Balasubramanian,
M.; Ren, Y.; Greeley, J. P.; Fister, T. T. J. Phys. Chem. Lett. 2011, 2,
2483.

AD

dx.doi.org/10.1021/cr400573b | Chem. Rev. XXXX, XXX, XXXXXX

Das könnte Ihnen auch gefallen