Vous êtes sur la page 1sur 12

REVIEW ARTICLE

PUBLISHED ONLINE: 15 DECEMBER 2011|DOI: 10.1038/NMAT3191

LiO2 and LiS batteries with high energy storage


Peter G. Bruce1*, Stefan A. Freunberger1, Laurence J. Hardwick1 and Jean-Marie Tarascon2
Li-ion batteries have transformed portable electronics and will play a key role in the electrification of transport. However, the
highest energy storage possible for Li-ion batteries is insufficient for the long-term needs of society, for example, extendedrange electric vehicles. To go beyond the horizon of Li-ion batteries is a formidable challenge; there are few options. Here we
consider two: Liair (O2) and LiS. The energy that can be stored in Liair (based on aqueous or non-aqueous electrolytes)
and LiS cells is compared with Li-ion; the operation of the cells is discussed, as are the significant hurdles that will have to be
overcome if such batteries are to succeed. Fundamental scientific advances in understanding the reactions occurring in the cells
as well as new materials are key to overcoming these obstacles. The potential benefits of Liair and LiS justify the continued
research effort that will be needed.

nergy storage will be more important in the future than at any


time in the past. Among the myriad energy-storage technologies, lithium batteries will play an increasingly important role
because of their high specific energy (energy per unit weight) and
energy density (energy per unit volume). Since their introduction in
1991, Li-ion batteries (Fig.1) have transformed portable electronic
devices14. New generations of such batteries will electrify transport
and find use in stationary electricity storage. However, even when
fully developed, the highest energy storage that Li-ion batteries can
deliver is too low to meet the demands of key markets, such as transport, in the long term. Reaching beyond the horizon of Li-ion batteries is a formidable challenge; it requires the exploration of new
chemistry, especially electrochemistry, and new materials2,5,6. There
are few options. Two, based on lithium, are receiving intense interest
at the present time and will be discussed in this Review: rechargeable
Liair (hereafter referred to as LiO2 as O2 is the fuel) and LiS batteries7. Other options, especially Znair, have been reviewed in detail
recently elsewhere814. Although LiO2 and LiS share the same
anode, and have active cathode components (O2 and S) that are nearest neighbours in group16 of the periodic table, there are important
differences related to the different chemistry of O and S and the different states of matter of their cathodes. LiS has been investigated since
the 1940s; the problems are formidable and extensive efforts have
been made to address them over the intervening 70years. Important
advances have made recently, but significant challenges remain7,1524.
In comparison, LiO2, especially with a non-aqueous electrolyte, has
received much less attention until recently 7,8,2530. As in the case of
LiS, major challenges will have to be solved if LiO2 batteries are
to succeed. The renaissance of interest in LiS and the upsurge of
interest in LiO2, based on aqueous and non-aqueous electrolytes,
reflects the need for electrochemical energy-storage devices that can
offer a leap forward; for example, delivering electric vehicles with a
driving range approaching the goal of ~500km between charging. In
the limited space available, we cannot hope to review all the excellent
work that has taken place on these two battery technologies. Instead,
we shall begin by considering the energy that can be stored in LiO2
and LiS cells, and then examine each system, how it operates, and
the challenges facing research that attempts to advance LiO2 and
LiS batteries. Paramount among the present challenges is a fundamental understanding of the chemistry taking place in the cells and
the discovery of new materials.

Energy storage from theory to practice

The theoretical specific energies (gravimetric energy densities) and


energy densities (volumetric energy densities) for LiS and LiO2
are given in Table1, where they are compared with those for Li-ion
and Znair. The values are based on the cell reactions in column 1,
that is, the energy obtained per unit mass or per unit volume of
the active components of the anode and cathode. Often a value of
11,586Whkg1 is quoted for the LiO2 cell; however this is based on
the mass of Li alone. All metalair cells gain mass (O2) as they discharge, so the mass of O2 should be included, as it is in Table1. The
leap forward in theoretical specific energy on migrating from Li-ion
to LiS and then LiO2 is clear. It arises because Li2S, Li2O2 and
LiOH in the cathode store more Li, and hence charge, than LiCoO2
per unit mass, and Li metal stores more charge per unit mass than a
graphite (C6Li) anode. The theoretical energy density is also greater
for LiO2 and LiS than Li-ion but the gain is not as great as for
specific energy.
Of course there is always a significant reduction in the energy
stored in a battery on moving from theory to practice. A comparison of practical specific energies for several rechargeable batteries
is presented in Fig. 2. The values for established technologies are
well attested but for LiO2 are at best very rough estimates, because
so far there are few realistic prototypes on which to base such figures. However, the values quoted are in line with those reported by
others16,26,28. Several factors conspire to lower the energy storage of
practical LiS and LiO2 batteries. The cathode in each case consists of a porous conducting matrix (usually carbon) in which the
discharge products form, thus adding mass and volume to the cell.
More Li metal than is required for the stoichiometric reaction has to
be included, to compensate for the inefficiency of Li-metal cycling.
The effect of these factors is illustrated in Box1 for the case of nonaqueous LiO2. Furthermore, packaging, current collectors and, in
the case of LiO2 cells, gas diffusion channels, will also reduce the
practical energy storage. There is a general rule of thumb that moving from theory to practice involves a reduction by a factor of 3
in energy storage. Prototype LiS cells have been developed by, for
example, Sion Power and, based on their data, the specific energy
of a LiS cell is reduced from 2,567Whkg1 to 350Whkg1, that
is, a factor of 7. However they state that values in the region of
600Whkg1 are expected in the near future (factor of 4) and similar
values have been suggested by others13,16,3133. If we apply a factor

School of Chemistry, University of St Andrews, North Haugh, St Andrews, Fife KY16 9ST, Scotland, UK, 2Laboratoire de Ractivit et Chimie des
Solides UMR CNRS 6007, 33rue Saint-Leu, 80039 Amiens Cedex, France. Present address: Stephenson Institute for Renewable Energy, Department
of Chemistry, University of Liverpool, Crown Street, Liverpool L69 7ZD, UK. *e-mail: pgb1@st-andrews.ac.uk

NATURE MATERIALS | VOL 11 | JANUARY 2012 | www.nature.com/naturematerials

2012 Macmillan Publishers Limited. All rights reserved.

19

REVIEW ARTICLE

NATURE MATERIALS DOI: 10.1038/NMAT3191

Li-ion
Discharge

LiO2 (non-aq)

LiO2(aq)

Discharge

Discharge

Li+
LixC6

Organic Li1-x CoO2


electrolyte

C6Li + Li0.5CoO2  3C + LiCoO2

LiS
Discharge
e

Li
metal

+
O2

Li
metal

Li

O2

Organic
electrolyte

2Li+ + 2e + O2

Porous
carbon +
catalyst

Li2O2
+

Li -conducting
membrane

Li2O2

+
O2
+

Li

Aqueous
electrolyte

O2
Porous
carbon +
catalyst

LiOH

Li
metal

Li+
Organic
electrolyte
Li2S

O2-evolution electrode

2Li+ + 2e + O2 + H2O

2LiOH

Porous
carbon
+S

2Li+ + 2e + S

Li2S

Figure 1 | Schematic representations of Li-ion, non-aqueous and aqueous LiO2 and LiS cells.

of 47 to the LiO2 battery then the estimated practical specific


energy is in the range of ~500900Whkg1, that is, the range given
in Fig.2, which is at least 23 times greater than Li-ion and would
be sufficient to deliver a driving range of more than 550km, when
scaled to the driving range of the Nissan Leaf (see caption to Fig.2).
Practical energy densities for LiO2 are even more difficult to estimate with any reasonable accuracy in the absence of realistic prototypes and may not exceed Li-ion, that is, the main gain over Li-ion
is in specific energy.
The practical specific energies for LiS and LiO2 given in Fig.2
represent the performances expected if the problems described in
the subsequent sections of this Review can be addressed successfully. The main factor limiting the practical energy storage of LiO2
and LiS cells is the need for excess Li in the anode; this especially
compromises volumetric energy density owing to the low density
of Li metal (0.534gcm3). Thus, improving the cycling efficiency of
the Li electrode or replacing it with an alternative (see later) is one of
the important challenges of maximizing the energy storage of LiO2
and LiS batteries.
The price of LiS and LiO2 cells will be as important as
performance in determining their adoption. The prices given in
Fig.2 are the targets at the pack level outlined by the US Advanced
Battery Consortium. Cells must be assembled into packs then into
systems that include the battery management. The US Advanced
Battery Consortium states that the target long-term minimal selling price of a mass-produced (25,000 units) 40KWh battery pack is
US$150kWh1. Clearly, whichever technology is used, a significant
reduction in cost is required to enable mass commercialization. Today
the Li-ion pack-level price is close to US$500600kWh1 (ref. 34).

The LiO2 battery

Schematic representations of LiO2 cells based on aqueous and


non-aqueous electrolytes are shown in Fig.1. In both cases, on discharge, the Li-metal anode is oxidized, releasing Li+ into the electrolyte, and the process is reversed on charge. At the positive electrode,
O2 from the atmosphere enters the porous cathode, dissolves in the
electrolyte within the pores and is reduced at the electrode surface
on discharge. When a suitable non-aqueous electrolyte is employed,
O22 is formed, which, along with Li+ from the electrolyte, forms
Li2O2 as the final discharge product. The peroxide is then decomposed on charging: 2Li++O2+2eLi2O2. Note that some authors
have reported that discharge down to Li2O is possible, which would
increase the energy stored (twice the Li per O, see Box1), but may
be difficult to reverse on charging 3538. Aqueous electrolytes involve
the formation of OH and then LiOH at the cathode on discharge,
according to the equation: 2Li++O2+H2O+2e2LiOH, with
20

LiOH being oxidized on charge. Because H2O and O2 are involved,


this is sometimes referred to as a Liwater battery. Note that we consider only LiO2 cells with alkaline electrolytes here, as these have
been more extensively studied than acidic electrolytes3941.
Although both cells involve O2 reduction on discharge, there
are important differences, especially relating to the reactions at the
cathode and the role of the electrolyte. As a result, aqueous and nonaqueous LiO2 will first be considered separately; the challenges
that exhibit some commonality with LiS, especially the Li anode
and porous carbon cathode, will then be discussed.

The non-aqueous LiO2 cell. The rechargeable non-aqueous LiO2


cell was first reported by Abraham and co-workers25 and interest
has expanded rapidly in recent years36,37,4259. Although important
progress has been made by many authors, significant challenges
remain; these are summarized in Fig.3.
The cathode, if exposed to ambient air, would be affected by CO2
and H2O, resulting in the formation of Li2CO3 and LiOH instead
of Li2O2. Thus it is necessary to remove these gases by, for example, an O2 diffusion membrane covering the outer surface of the
cathode that blocks CO2 and H2O. Such a membrane would have
to support fast O2 diffusion so as not to limit the rate. Zhang and
co-workers have examined many aspects of the non-aqueous LiO2
cell6063. These authors have investigated membranes to see if they
are suitable for protecting the non-aqueous LiO2 cell from ingress
of CO2 and H2O when operating in ambient air. They investigated
hydrophobic polymethylsiloxane and silicalite on a porous metal
substrate56,57, and Melinex (DuPont), a polyesterpolethyleneglycol
copolymer, or high-density polyethylene films62. Using such membranes it was possible to discharge the cell continuously for one
month in ambient air with 20% relative humidity, compared with
only a few hours in the same environment in the absence of the
membrane, demonstrating progress in using the cells in air.
The electrolyte is a key component and one of the main challenges
at present. It must be stable both to O2 and its reduced species, as
well as the LiOx compounds that form on discharge; it must exhibit
sufficient Li+ conductivity, O2 solubility and diffusion to ensure
satisfactory rate capability, as well as wet the electrode surface and
possess low volatility to avoid evaporation at the cathode.
Electrolytes based on organic carbonates (for example, LiPF6 in
propylene carbonate) have so far been widely used in non-aqueous
LiO2 cells, because of their low volatility, compatibility with Limetal and high oxidation stability (>4.5 V compared with Li+ or
Li). However, studies of O2 as a contaminant in Li-ion battery electrolytes showed that organic carbonates are susceptible to nucleophilic attack by reduced O2 species64. Recent studies of LiO2 cells
NATURE MATERIALS | VOL 11 | JANUARY 2012 | www.nature.com/naturematerials

2012 Macmillan Publishers Limited. All rights reserved.

REVIEW ARTICLE

NATURE MATERIALS DOI: 10.1038/NMAT3191


Table1 | Data for several electrochemical reactions that form the basis of energy-storage devices.
Battery

Cell voltage (V)

Theoretical specific energy (Wh kg1)

Theoretical energy density (Wh l1)

Todays Li-ion
C6Li+Li0.5CoO23C+LiCoO2

3.8

387

1,015

Znair
Zn+O2ZnO

1.65

1,086

6,091* (ZnO)

LiS
2Li+SLi2S

2.2

2,567

2,199 (Li+Li2S)

LiO2 (non-aqueous)
2Li+O2Li2O2

3.0

3,505

3,436 (Li+Li2O2)

LiO2 (aqueous)
2Li+O2+H2O2LiOH

3.2

3,582

2,234|| (Li+H2O+LiOH)

*Based on volume of ZnO at the end of discharge; based on the sum of the volumes of Li at the beginning and Li2S at the end of discharge; based on the sum of the volumes of Li at the beginning and Li2O2 at the
end of discharge; assuming the product is anhydrous LiOH and alkaline conditions; and ||based on the sum of the volumes of Li+H2O consumed and the LiOH at the end of discharge.

carbonates, which we now know decompose on discharge, as


discussed above65,66,69. Therefore it is difficult to know the extent to
which the catalyst is catalysing the electrolyte decomposition on discharge and the oxidation of the decomposition products on charge,
rather than Li2O2 formation and decomposition. The effect of the
catalyst on charging Li2O2 has been investigated by charging electrodes constructed with Li2O2 incorporated in the as-prepared cathode (Fig.4.)37,77. Even when organic carbonate electrolytes are used,
it has been shown by differential electrochemical mass spectrometry that only oxidation of Li2O2 occurs73. These studies were further
facilitated by the development of a fast screening technique based on
the catalytic degradation of H2O2 (H2O2(aq)H2O+O2), which
exhibits similar trends to the electrochemical oxidation of Li2O2
(Li2O2(s)2Li+O2). Investigation included a variety of transition
metal oxides, with nanowires of -MnO2 giving the lowest charging
voltage in LiO2 cells and the most facile H2O2 decomposition of the
materials investigated in the study 77. Overall, it has been demonstrated that catalysts lower the charging voltage of Li2O237,77.
Studies of O2 reduction in non-aqueous electrolytes have
been carried out for several decades in solvents more stable than
organic carbonates64,78. However, it is important to understand the

1,000
Specific energy (Wh kg1)

with organic carbonate electrolytes have demonstrated electrolyte


degradation does indeed occur on discharge6569. Significantly, studies have also shown that there is little or no evidence of Li2O2 formation occurring in parallel with the electrolyte degradation69. This is
important as it rules out the possibility that the remarkable ability
of LiO2 cells with organic carbonate electrolytes to sustain cycling
(up to 100 cycles65) could be explained by reversible Li2O2 formation, whereas the ubiquitous capacity fading (Fig. 3) is explained
by simultaneous electrolyte degradation. Instead, such cells cycle
by degradation of the electrolyte on discharge to form lithium
propyl dicarbonate (C3H6(OCO2Li)2), Li2CO3, HCO2Li, CH3CO2Li,
CO2 and H2O, with the decomposition products being oxidized on
charge69. The implication of these results is that all studies of the
LiO2 cell with organic carbonate electrolytes are likely to be significantly affected by electrolyte degradation on discharge. This highlights the importance of presenting data from analytical methods
(for example, Fourier transform infrared, Raman and mass spectrometry) to demonstrate that the product actually formed on discharge is Li2O2, when reporting results on LiO2 cells.
Recently, attention has turned to other electrolytes, especially
ethers, including tetraglyme (CH3O(CH2CH2O)4CH3), dimethoxyethane (CH3OCH2CH2OCH3; DME) and polyethyleneoxide
(PEO). Such ethers are certainly more stable than organic carbonates towards reduced O2 and do exhibit Li2O2 formation at ~2.7V, at
least on the first discharge42,43,64,7072. Recently, cells using DME and
LiClO4 electrolyte and all-carbon-nanofibre electrodes have shown
enhanced cycling stability compared with typical cells with carbonate-based electrolyte, though fading is still present 59. However,
investigation of linear (diglyme, triglyme and tetraglyme) and cyclic
(1,3-dioxolane, 2-methyl tetrahydro furan) ethers in LiO2 cells
demonstrated electrolyte degradation to form Li2CO3 and HCO2Li,
and CH3CO2Li, which becomes more severe on cycling 70,71. Studies
of DME in LiO2 cells by mass spectrometry showed that only ~60%
of the O2 consumed on discharge is released on charge60. If these
results are confirmed by further studies then the implication is that
ethers, although more stable than carbonate electrolytes, are not the
final solution to the challenge of identifying a suitable non-aqueous
electrolyte for LiO2 batteries.
Provided Li2O2 can be formed on discharge then it has been
shown that it can be oxidized on charge49,73. Such charging has been
investigated by constructing electrodes in the discharged state, that
is, by incorporating Li2O2 in the electrode49,73. The dominant gas
evolved on charging is O2, consistent with Li2O2 oxidation without
electrolyte decomposition71,73.
The gap between the charge and discharge voltage has been
linked to the nature of the catalyst in the positive electrode, with
values ranging from 0.6 to 1.5 V37,46,4951,7476. However, many of
these studies were carried out in electrolytes containing organic

> 550 km

800
> 400 km

600
> 200 km

160 km

200
0

> 225 km

Today

400

50 km 80 km 100 km
Pbacid NiCd NiMH Liion Future Znair LiS
Liion

Price (US$ kW h1) 200

600

900

Available

600

Liair

< 150 < 150

< 150

Under development

R&D

< 150

Figure 2 | Practical specific energies for some rechargeable batteries,


along with estimated driving distances and pack prices. For future
technologies, a range of anticipated specific energies are given as shown by
the lighter shaded region on the bars in the chart for rechargeable batteries
under development and R&D. The values for driving ranges are based
on the minimum specific energy for each technology and scaled on the
specific energy of the Li-ion cells (140Whkg1) and driving range (160km)
of the Nissan Leaf136. The prices for technologies under development
represent targets set by the US Advanced Battery Consortium137.

NATURE MATERIALS | VOL 11 | JANUARY 2012 | www.nature.com/naturematerials

2012 Macmillan Publishers Limited. All rights reserved.

21

REVIEW ARTICLE

NATURE MATERIALS DOI: 10.1038/NMAT3191

Box 1 | Factors influencing the energy storage of a practical LiO2 battery.

Excess Li

Beginning with the simplest model for a non-aqueous LiO2 cell,


assuming the cathode is composed only of Li2O2, (that is, no porosity, or carbon, or binder) and assuming a stoichiometric quantity
of Li in the anode, see Fig. B1a; the specific energy and energy
density are 3,505Whkg1 and 3,436Whl1 respectively, that is, the
values given in Table1. The effect of introducing excess Li metal
in the anode to compensate for the Li loss on cycling is shown. For
a threefold excess (n=4) the specific energy and energy density
become 1,800Whkg1 and 1,290Whl1. Values for discharge to
Li2O are also shown.

Li2O2

3,500
3,000
2,500
2,000
1,500
1,000

2
3
Equivalents of Li

or

Li + O2

Li2O

3,500
3,000
2,500
2,000
1,500
1,000

5,000

5,000

4,000

4,000

3,000

3,000

2,000

2,000

1,000

2
3
Equivalents of Li

1,000

Specific
energy (Wh kg1)

Li2O

Energy
density (Wh l1)

Li consumed

Li + O2

Taking account of porosity in the cathode, let us assume 20% of


the cathode volume is occupied by C, 20% by electrolyte and 60%
by Li2O2 at the end of discharge. The resulting specific energies
and energy densities are shown in Fig. B1b, where the porous
cathode is coupled with a Li anode; a stoichiometric amount of Li
is assumed to provide a benchmark for the highest possible energy
storage that is, in principle, possible. Comparison with a Li-ion
cell is shown. Bars framed green represent values for Li2O2, those
framed orange are for discharge to Li2O. The electrolyte thickness
is 10 m.

Specific
energy (Wh kg1)

Li2O2

Energy
density (Wh l1)

Porous cathode

b
4,000

Maximum LiO2

4,000
Maximum LiO2

3,000
Li2O2
2,000

Volumetric
energy density (Wh l1)

Gravimetric
specific energy (Wh kg1)

Li2O

Li2O2
2,000

1,000

1,000

Li2O

3,000

LiCoO2/C

LiCoO2/C
0

Figure B1 | Specific energy and energy densities of the non-aqueous LiO2 battery. a, Excess Li in the anode. b, A porous cathode. The values for
todays Li-ion battery (LiCoO2/C) shown as comparison.

fundamental mechanism of O2 reduction in the presence of Li+ and


the formation of Li2O2, as well as the mechanism of Li2O2 oxidation on charging, if progress is to be made on LiO2 cells. Detailed
electrochemical investigations have probed the influence of salt and
solvent on O2 reduction36,42,72. Insitu spectroelectrochemical measurements involving Raman and mass spectrometry have also been
applied, which offer the important benefit of directly identifying the
species involved in the reactions (intermediates and products) on
discharge and charge73. (Note that in ref.73 the peak maximum for
O2 reduction in the cyclic voltammogram is 2.2V, however the peak
22

onset (the start of the reduction) is at 2.7V, corresponding to the


voltage of the discharge plateau.)
The consensus from the various studies is that the mechanism of
O2 reduction on discharge is:
O2+eO2

(1a)

O2+Li+LiO2

(1b)

2LiO2Li2O2+O2

(1c)

NATURE MATERIALS | VOL 11 | JANUARY 2012 | www.nature.com/naturematerials

2012 Macmillan Publishers Limited. All rights reserved.

REVIEW ARTICLE

NATURE MATERIALS DOI: 10.1038/NMAT3191


Anode

Cathode

Discharge
e

e
Li metal

Li

Capacity (mAh g1 C)

1,000

10

15

20

CO2

Li2O2

Electrolyte
Stability
Conductivity
Volatility
O2 solubility, diffusivity

Charge

Porous
carbon +
catalyst

Voltage gap

4.5

Discharge

H2O

O2

Capacity fading

O2

O2

Li
metal

Organic
electrolyte

2,000

Cathode needs a membrane to block CO2


and H2O, while allowing O2 to pass.

25

Potential (V versus Li/Li+)

Problems of Li metal
Dendrite formation
Cycling efficiency
Requires stable solidelectrolyte interphase
Safety issues

Charge

4.0
3.5
3.0
2.5
Discharge

2.0

30

1,000

2,000

Capacity (mAh g1 C)

Cycle number

Porous cathode design


Pore size, distribution
Catalyst type, distribution, loading

Figure 3 | Challenges facing the non-aqueous LiO2 battery. Data illustrating capacity fading and the voltage gap were collected from a cell with an organic
carbonate electrolyte, Li1MLiPF6 in propylene carbonate(superP:Kynar:-MnO2 nanowires)69. Use of more stable ether electrolytes and carbon fibre
nanotube electrodes for example, LiLiClO4 in dimethoxyethanecarbon fibre nanotube reduces, although does not eliminate, capacity fading59.

LiO2+Li++eLi2O2

(1d)

Li2O2+Li++e2Li2O

(1e)

Equation (1d) occurs at lower voltages than equation (1a)73 and


so far there has been limited evidence for the formation of Li2O.
Oxidation of Li2O2 follows:
Li2O22Li++2e+O2

(2)

In other words, the process on charging is not the reverse of


discharge; the latter involves O2 as an intermediate, whereas the
former does not. The different pathways followed on reduction and
oxidation do not violate the principle of microscopic reversibility,
but arise because the kinetics of oxidizing Li2O2 directly are faster
than reversing the three steps on reduction, especially disproportionation73. The result of these different pathways is the observed
separation of the charge and discharge voltages. Another factor
that may contribute to the voltage gap is if singlet O2 is formed
on oxidation of Li2O2, whereas reduction involves the more stable
triplet state72. The singlet-to-triplet O2 transition is spin forbidden
and hence kinetically hindered. If the transition kinetics are significantly slower than O2 evolution on oxidation of Li2O2, then the
difference in energy between the singlet and triplet states (~0.9V)
could influence the voltage gap. Because voltage gaps smaller than
0.9V have been observed (typically 0.7V), at most any difference
in the spin states of O2 is likely to make only a partial contribution
to the gap.
Other electrolytes have been explored, but in much less detail
than the organic electrolytes. Investigation of hydrophobic ionic
liquids demonstrated that they can maintain less than 1% H2O

content after 100hours of operation, in the case of 1-ethyl-3-methyl


imidizolium bis(trifluoromethane sulphonyl)imide. On discharge,
a capacity of 5,360 mAh g1 (based on carbon alone) has been
reported for operation of up to 56days52. Solid electrolytes have also
been investigated, in particular, cells incorporating the Li+ conductor 18.5Li2O:6.07Al2O3:37.05GeO2:37.05P2O5 (LAGP) can sustain
some 40 cycles at elevated (40100C) temperatures79. Cells with a
solid PEO-based polymer electrolyte can be charged at relatively low
voltages, as noted above72. However, in view of what is now understood about the reactions in liquid electrolytes, it will be important
in the future to investigate the nature of the discharge products and
the electrode reaction mechanisms in the ionic liquid and solid
4.5
e 2O 3
-F

No catalyst
Potential (V versus Li/Li+)

Several studies suggest also the direct reductions36,72:

NiO
CuO

4.0
Co3O4

Electrolytic MnO2
-MnO2 bulk

3.5

3.0

250

500

-MnO2 nanowire

750

1,000

1,250

Capacity (mAh g1 C)

Figure 4 | First galvanostatic charge, i=70mAg1 C (that is, Li2O2


oxidation) for various catalyst-containing LiO2 cells in this study77.
Figure adapted with permission from ref.77, 2010 ECS.

NATURE MATERIALS | VOL 11 | JANUARY 2012 | www.nature.com/naturematerials

2012 Macmillan Publishers Limited. All rights reserved.

23

REVIEW ARTICLE

NATURE MATERIALS DOI: 10.1038/NMAT3191


Discharge

Anode
Problems of Li metal
Cycling efficiency
Li/ceramic interface
Safety issues

Cathode
Cathode needs membrane
blocking CO2 , while allowing
O2 to pass

e
e

+
O2

Li
metal

Requires Li+conducting
membrane (such as
LISICON) to protect Li.
Precipitation of LiOH on
LISICON membrane and
membrane instability in
strong acidic and
basic electrolyte.

Li+-conducting
membrane

Li+

O2

Aqueous
electrolyte

Porous
carbon +
catalyst

LiOH

O2-evolution electrode
Requires cheap, efficient
O2evolution catalyst.
Electrolyte
Limited LiOH solubility in H2O:
precipitation of LiOH, blocking
electrode pores.

O2
CO2
Optimize gas-diffusion
electrode for LiO2 cell.
O2 + H2O 2OH

Requires cheap, efficient


oxygen-reduction catalyst.

Figure 5 | Challenges facing the aqueous LiO2 battery. Image of LISICON membrane reproduced with permission from ref.39, 2010 ECS.

electrolytes before any conclusions can be formed concerning their


potential use in practical LiO2 cells.

The aqueous LiO2 cell. Pioneering work by Visco and colleagues


in 200726, involving the development of protected Li anodes, opened
the way to the rechargeable aqueous LiO2 battery 26,40,41. Additional
challenges facing rechargeable aqueous LiO2 are summarized
in Fig.5.
Unsurprisingly, significant attention has focused on protecting the
Li-metal anode from the aqueous electrolyte, using a Li+-conducting
but electronically insulating membrane, such as LISICON-type glass
ceramic (lithium superionic conductor, Li(1+x+y)AlxTi2xSiyP(3y)O12),
without which the rechargeable aqueous LiO2 cell could not function. This is discussed later in the section comparing LiO2 and LiS
as it also has applications in non-aqueous LiO2 and LiS.
The Li anode is far from the only problem facing aqueous LiO2
cells. Although one of the advantages of such a cell is that there is,
by definition, no need to avoid ingress of H2O from the atmosphere,
CO2 remains a problem as it leads to the formation of Li2CO3. As a
result, it is necessary to apply a protective membrane to the outer
5.0

Charge

4.5
4.0

Cell voltage (V)

3.5
3.0

Discharge

2.5
2.0
1.5
1.0
0.5
0.0
0

20

40

60

80

100

120

140

160

Capacity (mAh g-1)

Figure 6 | Load curve of an aqueous LiO2 cell. Cell cycled over a limited
range to avoid precipitation of LiOH. Figure reproduced with permission
from ref.41, 2010 ECS.
24

surface of the cathode that permits O2 diffusion while blocking CO2,


or to remove the latter from the atmosphere around the battery by
alternative means.
The aqueous electrolyte is not inert, it takes part in the cathode
reaction, see Table 1. As discharge proceeds, the solvent is consumed and LiOH is generated, resulting in a rapid increase in the
concentration of the latter. Saturation is reached at 5.3Ml1, corresponding to a specific capacity of only ~170mAhg1 (based on
the electrolyte alone) that is, barely more than a Li-ion cell. To avoid
limiting the capacity, a flow-cell design, in which the electrolyte is
constantly replenished with a fresh solution, has been proposed; the
same authors have also proposed recharging by collecting the LiOH,
extracting the Li metal then mechanically replacing the anode80,81.
Otherwise, to achieve a specific capacity that offers a meaningful
advantage over Li-ion cells, LiOH must be allowed to precipitate
as a solid. Such precipitation has important consequences; not only
will solid LiOH block the ceramic protecting the anode, but it can
also clog the porous cathode. One solution to the latter is the use of
an anion-exchange membrane at the cathode/electrolyte interface;
this permits OH generated in the porous cathode to be transported
out, while blocking Li+ entering the electrode, hence LiOH precipitates outside the cathode. This membrane also blocks any carbonates that might be formed by CO2 ingress, preventing the formation
of Li2CO3. An interesting solution to the problem of recharging the
aqueous LiO2 cell containing solid LiOH is the incorporation of a
third electrode, which is used for O2 evolution on recharging (Fig.5).
The third electrode, which remains immersed in the liquid aqueous
electrolyte, was found to more effectively oxidize LiOH than if solid
LiOH is allowed to precipitate in the pores of the carbon cathode.
Such a cell has been demonstrated to operate for several hundreds
of cycles39. As in the case of the non-aqueous LiO2 cell, a voltage
gap between charge and discharge is evident. At a current density of
0.1mAcm2 this can be as low as 0.3V (ref.39) or ~0.9V at a higher
current density of 0.5mAcm2 (Fig.6)41.
The reduction and evolution of O2 in aqueous media have been
studied extensively for many years. In contrast to the reduction of
O2 to Li2O2 in non-aqueous electrolytes, reduction to LiOH necessitates cleavage of the OO bond, a reaction that requires a catalyst, as
does O2 evolution. The most widely used catalyst is Pt (refs 39,41),
although other materials, including manganese oxides82, metallic
copper 83 and various perovskite oxides84 have been investigated.
One advantage of separating out the electrodes for O2 reduction
NATURE MATERIALS | VOL 11 | JANUARY 2012 | www.nature.com/naturematerials

2012 Macmillan Publishers Limited. All rights reserved.

REVIEW ARTICLE

NATURE MATERIALS DOI: 10.1038/NMAT3191


Discharge

Anode

Li
metal

Li+
Organic
electrolyte
Li2S

Porous
carbon
+S

Potential (V versus Li/Li+)

Capacity fading

Discharge capacity (mAh g1)

S8

e
Li metal

Problems of Li metal
Dendrite formation
Cycling efficiency
Requires stable solidelectrolyte interphase
Safety issues
Formation of insulating Li2S
layer on Li anode

Cathode

Address
polysulphide solubility
Charge process

3.0

1.5

Li2S8

Li2S6
Li2S4
Discharge process
Capacity (mAh g1)

Li2S2
Li2S
1,000

1,600
Electrolyte
Stability
Conductivity
Solubility of polysulphides

1,200
800

Trap discharge
products in mesoporous
carbon cathode.
3 nm
6.5 nm

400
0

10

20
Cycle number

30

40

Figure 7 | Challenges facing the LiS battery. Load curve and schematic showing PEG200-coated CMK-3S composites that impede diffusion of the
polysulphides into the electrolyte; reproduced from ref.17, 2009 NPG. Capacity fading for LiS cell using graphene-nanosheets cathode; adapted with
permission from ref.21, 2011 Elsevier.

and evolution is that the former is not degraded by O2 evolution


and, in principle, different catalysts can be employed at the different
electrodes, avoiding the need for both catalysts to be stable in the
same voltage range.
Although the basic mechanisms of O2 reduction and evolution in
aqueous electrolytes are well known, the specific processes in those
containing lithium salts have received much less attention and merit
further study in view of the present interest in aqueous LiO2 cells84.
In this Review we have focused on alkaline electrolytes, as these
have been most widely used in LiO2 cells so far. However, acidic
electrolytes may also be used41 and these give rise to higher voltages
(up to ~4.25V versus Li/Li+): 2Li+O2+2H+2Li++H2O.

The LiS battery

The rechargeable LiS cell is shown in Fig. 1 and operates by


reduction of S at the cathode on discharge to form various polysulphides that combine with Li to ultimately produce Li2S. Such
cells have many attractive features, including: (i) the natural abundance and low cost of S; and (ii) high theoretical energy storage
(Table1)1517. Yet the promise of a device with greater energy storage
and cycle life than Li-ion has not yet materialized; even after decades
of development, the LiS battery has still not reached mass commercialization. Several problems inherent in the cell chemistry remain
and are summarized in Fig.7. Among such problems, discussed in
detail in a previous review 16, are: (i) poor electrode rechargeability and limited rate capability 85,86 owing to the insulating nature of
sulphur and the solid reduction products (Li2S and Li2S2); (ii) fast
capacity fading owing to the generation of various soluble polysulphide Li2Sn (3n6) intermediates8790, which gives rise to a shuttle
mechanism91; and (iii) a poorly controlled Li/electrolyte interface.
The shuttle mechanism arises because the soluble polysulphides
that are formed at the cathode are transported to the anode where
they are reduced to lower polysulphides, which are then transported

back to the cathode, where they become reoxidized and then return
to the anode. If, at the anode, reduction proceeds to form insoluble
Li2S2 or Li2S, then this can deposit on the anode and elsewhere.
Much of the recent work to improve LiS cells builds on previous
approaches. Considerable effort has been devoted to designing
porous composite cathodes that are capable of delivering electrons
efficiently to the S as well as trapping the soluble polysulphides.
These aspects of LiS battery research are described in the section
comparing LiO2 and LiS.
A different approach to the problem of minimizing transport of
the soluble polysulphides from cathode to anode involves the use
of organosulphur-based polymer systems with SS linkages offering high specific energy and being capable of reversibly cleaving
and reforming on reduction and oxidation in the molecular skeleton9294. The charge/discharge reactions of the systems are based
on the redox chemistry of thiolates (RS), which can be oxidized
to give the corresponding radical (RS), which can, in turn, couple
to form disulphides (RSSR)95,96. Poly(2,2-aminophenyl-disulphide)
was an early example96. The redox chemistry of the dimercaptothiadiazole polymer has also been extensively studied and was shown to
polymerize to form the highly insoluble polydisulphide. A marked
advantage of conjugation with the electron-poor thiadiazole ring is
a substantial increase in the discharge potential plateau by approximately 0.62.8 V. Asides from this, polyvinyl disulphide polymers were also shown to deliver a sustainable reversible capacity
of 400 mAh g1 for at least 200 cycles97, and other polyvinyl sulphides containing more sulphur atoms in Sn units (2<n<7) are
being studied at present. Last, it should be recalled that earlier,
Degott reported92 polythiene-type conjugated polymers with the
expected large capacity of 630mAhg1. However, the polymer had
poor kinetics owing to the undesirable crosslinks between chains.
Incorporating this approach with mesoporous carbon electrodes
may offer an interesting way forward.

NATURE MATERIALS | VOL 11 | JANUARY 2012 | www.nature.com/naturematerials

2012 Macmillan Publishers Limited. All rights reserved.

25

REVIEW ARTICLE

NATURE MATERIALS DOI: 10.1038/NMAT3191

As for LiO2, the Li-electrolyte interface constitutes another


important challenge for the LiS battery that is discussed in the
section comparing LiO2 and LiS.
Finding an electrolyte to combat the irreversible loss of sulphur
associated with the formation of soluble Li2S6, Li2S4 and insoluble
Li2S2 or Li2S is one of the greatest challenges for LiS. Yet solutions remain elusive, even after intensive study for many years.
Special attention has been given to cyclic or linear ethers such as
tetrahydrofuran87,98101 1,3-dioxolane23,99101, dimethoxyethane23 and
tetra(ethylene glycol)-dimethyl ether 102 because of the absence of
chemical attack from S centres. However, certain general trends
have emerged: (i)solvents with high permittivity and donor numbers solvate long- and short-chain lithium polysulphides; and
(ii)solvents with high viscosity tend to inhibit the dissociation of
S82 and disproportionation of lithium polysulphide.
Such differences explain the strong dependency of the shape of
the voltagecapacity curve (for example, length of the first versus
second plateaus) on the nature of the electrolyte. It is unlikely that
a single solvent can fulfil all the requirements of an electrolyte for
the LiS battery, based on the long-standing research on mixtures
of solvents103,104 or even the exploration of ionic liquids22. The
likelihood of a breakthrough in electrolyte design, although not
impossible, is slim unless fresh directions, such as reversible organic
or inorganic species that chelate/bind sulphide species, are explored.
Faced with the problem of liquid electrolytes, researchers have
investigated solid-state LiS batteries, using a wide variety of polymers (PEO-based)105 or gel (polyvinylidene-fluoride-based)20,106
electrolytes or inorganic glassceramic electrolytes (Li2SP2S5)
combined with a Cu and S positive electrode composite107. Success
was limited, with cells suffering from rapid capacity decay or large
polarization from the persistence of lithium dendrite formation and
polysulphide dissolution. These problems were not alleviated by
going to the solid state.
Many strategies have focused on one component of the cell. In
contrast, Scrosati and colleagues18 have adopted a holistic approach
by modifying the entire cell. Their cell is assembled in the discharged state (Li2S) by using a lithium sulphidecarbon composite
(made by ball milling) as the cathode. The Li negative electrode is
replaced by a SnCLi alloy with higher chemical stability towards
sulphides and not prone to dendrite formation. The electrolyte is a
polymer-gel membrane formed by gelling with 1MLiPF6 in ethylene carbonate and dimethyl carbonate, saturated with Li2S, with a
PEO and lithium trifluoromethane sulphonate (LiCF3SO3) polymer
matrix incorporating a ZrO2 nanofiller. The cell can sustain a specific
energy of ~1,100Whkg1 for tens of cycles, a value not previously

Specific capacity (mAh g1)

1,400
1,200
1,000
800
600
400
200
0

10
Cycle number

15

20

Figure 8 | Cycle-life data of LiS cell. Comparison of LiS cells with a


cathode composed of S imbibed into the mesoporous carbon CMK-3
incorporating PEG (upper points) versus a similar cell but with a
mechanical mixture of CMK-3 and S (lower points) at a rate of 168mAg1
and at room temperature. Figure reproduced from ref.17, 2009 NPG.
26

achieved for a Li-metal-free battery. The relatively large polarization


(~2V) between charge and discharge, which penalizes the energy
efficiency of the system, could be addressed by substituting Sn for
the Si electrode108.

Some common challenges of LiO2 and LiS

Although different in a number of ways, there are important


similarities between some of the challenges facing aqueous and
non-aqueous LiO2 as well as LiS, summarized in Figs3, 5 and 7.
All rely on porous positive electrodes. Several strategies have been
explored to develop cathodes with engineered porosity and structure capable of efficient electron transport to the S and capture of
the polysulphides formed on discharge; two of the most important
challenges for LiS cells. Most of the approaches, since the work of
Peled et al. in 198890, have involved the fabrication of disordered
mesoporous carbonsulphur composites86,109112 occasionally loaded
with inorganic nanofillers, for example, Mg0.6Ni0.4O (ref.113), Al2O3
(ref. 114), or composites with sulphur embedded in electronically
conducting polymers to immobilize polysulphides115117. However,
the most significant advance has been made by Nazar and colleagues17 who have demonstrated that cathodes based on ordered
nanostructured mesoporous carbonsulphur composites provide
higher and more sustained, reversible capacities (Fig.8). They also
demonstrated the feasibility of further delaying the diffusion of
polysulphide out of the cathode structure by functionalizing the
pore surfaces of the carbon with polyethylene glycol (PEG) chains of
varying molecular weights17. Subsequently, similar levels of performance were observed using porous carbonaceous composites118120.
Sandwich-type graphene-sheetsulphur nanocomposites, where the
expanded graphite or graphene layers act as sulphur microcontainers, exhibit excellent performance121,122. Overall, these improvements
were ascribed to the intimate contact on the nanoscale between the
insulating sulphur and the conductive carbon framework, while
maintaining accessibility for the electrolyte. A further approach
involving nanodimension sulphurpolythiophene coreshell or
nanosulphur(polypyrrole-co-aniline) composites123,124 exhibits
a similar cycle life to the ordered mesoporous carbons. A recent
advance that further stabilized the sulphur cathodes relies on the
concept of using polysulphide reservoirs composed of porous silica19
or metalorganic framework additives125 embedded in the carbon
sulphur composite. Such approaches reduce the capacity fading to
values of ~80% of the initial capacity after 80 cycles. This is a significant advance, yet is still not sufficient for most applications.
For LiS, the insulating S resides in the pores, which also trap
the LinS discharge products. For LiO2 the porous electrode must
support diffusion of O2 gas, so that O2 is transported to the electrolyte/electrode interface as much as possible through the gas phase
rather than by slower diffusion in the electrolyte. In the case of the
non-aqueous cell, clogging of the pores by the solid discharge product, Li2O2, must be minimized. A variety of carbon substrates has
been investigated by several authors43,49,54,62,76,126,127. Pores that are too
small lead to clogging whereas those that are too large may inhibit
recharging. Mesoporosity is expected to offer the best compromise
and be able to accommodate the catalyst particles, which must be
distributed across the pore surface. The deposition of an insulating solid such as Li2O2, if it formed a continuous film, would soon
block the electrode surface (its growth limited by electron tunnelling, assuming ionic conductivity is lower than electronic in Li2O2).
Controlling the morphology of the discharged products in the pores
is therefore important 128. Ensuring some solubility for Li2O2 is an
advantage as it would promote higher charge and discharge rates.
It is also important to ensure that the electrolyte wets the pore surfaces. Most porous electrodes are based on carbon, because of its
low cost, high conductivity and processability. Concerning the stability of the carbon, oxidation (corrosion) may occur if high voltages are accessed on charging. A recent study has indicated that
NATURE MATERIALS | VOL 11 | JANUARY 2012 | www.nature.com/naturematerials

2012 Macmillan Publishers Limited. All rights reserved.

REVIEW ARTICLE

NATURE MATERIALS DOI: 10.1038/NMAT3191


carbon is relatively stable on reduction (discharge) in non-aqueous
LiO2 cells71. In the case of aqueous cells, as noted above, the preferred three-electrode cell configuration should minimize deposition of solid LiOH in the porous cathode. The main challenge here
is therefore to design a good gas-diffusion electrode incorporating a
low-cost reduction catalyst.
The Li negative electrode, whether coupled with O2 or S as the
positive electrode, suffers from several problems (Figs3, 5 and 7).
The growth of mossy Li on charging requires excess Li to mitigate
against loss on each cycle, adding a severe penalty to the overall
energy density of the system. The limited amount of Li resources
leads to concerns about Li shortage in view of the increasing
demands of automotive transportation. The parasitic consumption of the metal can occur by oxidation from O2 crossover from
the porous cathode or by reaction with (poly)sulphides leading to
insoluble Li2S species at the negative electrode. Such issues must be
addressed if we are to develop Li-lean and long-lasting LiO2 or LiS
systems. A twin-track approach aimed at improving the Li electrode
and exploring alternative anodes to Li metal should be adopted.
Considering first improving the Li metal anode, neither the dynamics of the Li surface roughness on cycling nor the chemistry of the
solvent reactions with Li have been fully addressed, despite considerable effort involving: (i) new anode designs; (ii) new surface treatments; (iii) Li+-conducting barriers; and (iv) the use of additives.
Encouraging results were obtained by adding LiNO3 to the cathode
or the electrolyte in LiS cells129, which was shown to modify the
solid electrolyte interphase on Li, thus increasing the cycling stability 24. Undoubtedly LiO2 and LiS cells will benefit greatly from the
use of a Li+-conducting ceramic, such as LISICON, protecting the Li
metal as it isolates the negative electrode chemistry from that at the
positive electrode, hence widening the choice of solvents to be used,
as they are no longer exposed to Li metal. This, for instance, enables the use of dimethylformamide or dimethylsulphoxide for LiO2
cells. Additionally, the ceramic will suppress the shuttle mechanism
in LiS cells. The Li+-conducting ceramic protecting the Li anode
is relatively thick (minimum 0.3 mm in the case of LISICON)130
and adds mass and volume to the cell. It is therefore important to
develop membranes that are thinner. Furthermore, the LISICON
ceramic is unstable in contact with Li and cycling the Li/ceramic
interface is difficult. To solve these problems a Li+-conducting layer
(typically based on a polymer, gel or non-aqueous liquid electrolyte) is placed between the Li anode and the ceramic26,40,41,131133. The
ceramic/electrolyte interface also has problems. In the case of LiS
cells, LISICON is partially reduced by the polysulphides, hence it
is necessary to add a protective membrane or change the ceramic
to enhance its stability against sulphur reduction. Similarly the
ceramic/aqueous-electrolyte interface suffers from degradation at
the extreme ends of the pH scale in the electrolyte in aqueous Li
O2 cells, and can become obstructed by deposition of LiOH39. To
address this problem, a cation-exchange polymer has been applied
to the surface of the ceramic that is exposed to the aqueous electrolyte39. Alternatively the electrolyte may be modified, for example, by
addition of acetic acid41. Overall, these issues call for greater efforts
in developing Li+-conducting ceramic electrolytes with enhanced
chemical stability.
A different approach to protecting the Li metal to improve its
cyclability is to use a different negative electrode, for example, Si
or Sn, which operate through alloyingdealloying reactions. Such
electrodes would have to be prelithiated at some point in the cellassembly process, unless the cells could be assembled in the discharged state with Li2O2 or Li2S in the cathode in the case of LiO2
and LiS, respectively. However, for LiS, this has proved to be
problematic owing to the resulting poor reversibility and high
polarization134. Such approaches do not prevent possible oxidation
of the lithiated alloys by O2 crossover from the cathode in LiO2
cells. For the cathode, a single discharge of a LiSO2 hybrid was

recently demonstrated by Zhang and colleagues135. It is argued that


the presence of O2 in the cell could be beneficial as it would oxidize the soluble polysulphide into insoluble sulphur, thus reducing
self-discharge.

Summary

LiS and LiO2 cells both offer substantial increases in specific


energy compared with Li-ion, but the gain in energy density is, at
best, modest. LiO2 has a higher specific energy than LiS. O2 is
free and S is cheap, and the main constituent of the cathode in both
cases is C. All represent advantages in minimizing the cost of the
cells compared with Li-ion. LiS is closer to mass commercialization
than LiO2, but the problems of LiS have been known for many
years and have not yet been completely solved. LiO2 has received
much less attention than LiS until recently, which means, as yet,
no intractable problems have been identified, but of course such
problems may be discovered in the future. It is important that future
work focuses on understanding the fundamental chemistry taking
place in the LiO2 and LiS cells if progress is to be made; otherwise any attempt to develop a technology is likely to fail. The recent
observation that the operation of LiO2 cells with organic carbonate
electrolytes is dominated by decomposition reactions and not Li2O2
formation highlights the importance of fundamental studies. There
is a degree of commonality in some of the problems facing LiS and
LiO2, especially concerning the Li anode and the porous C cathode.
It is not yet clear if any of the advanced batteries that offer to
take us beyond Li-ion will become a commercial success. However,
despite the difficulties, our society needs energy-storage devices
with much higher levels of energy storage than ever before. LiO2
and LiS batteries are among the few contenders that can exceed the
stored energy of Li-ion. We must devote more intensive research to
address the problems of the LiO2 and LiS batteries.
Corrected after print 4 January 2012

References
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.

Nagaura, T. & Tozawa, K. Lithium ion rechargeable battery.


Prog. Batteries Sol. Cells 9, 209217 (1990).
Tarascon, J.M. & Armand, M. Issues and challenges facing rechargeable
lithium batteries. Nature 414, 359367 (2001).
Schalkwijk, W.v. & Scrosati, B. Advances in Lithium-Ion Batteries (Kluwer
Academic/Plenum, 2002).
Nazri, G-A. & Pistoia, G. Lithium Batteries: Science and Technology
(Springer, 2003).
Bruce, P.G. Energy storage beyond the horizon: Rechargeable lithium
batteries. Solid State Ionics 179, 752760 (2008).
Bruce, P.G., Scrosati, B. & Tarascon, J-M. Nanomaterials for rechargeable
lithium batteries. Angew. Chem. Int. Ed. 47, 29302946 (2008).
Bruce, P.G., Hardwick, L.J. & Abraham, K.M. Lithium-air and lithium-sulfur
batteries. Mater. Res. Soc. Bull. 36, 506512 (2011).
Lee, J-S. etal. Metalair batteries with high energy density: Liair versus
Znair. Adv. Energy Mater. 1, 3450 (2011).
Neburchilov, V., Wang, H.J., Martin, J.J. & Qu, W. A review on air cathodes
for zinc-air fuel cells. J.Power Sources 195, 12711291 (2010).
Li, Q.F. & Bjerrum, N.J. Aluminum as anode for energy storage and
conversion: a review. J.Power Sources 110, 110 (2002).
Beck, F. & Ruetschi, P. Rechargeable batteries with aqueous electrolytes.
Electrochim. Acta 45, 24672482 (2000).
Encyclopedia of Electrochemical Power Sources (Elsevier, 2009).
Hamlen, P. & Atwater, T.B. Handbook of Batteries (McGraw-Hill, 2001).
Duduta, M. etal. Semi-solid lithium rechargeable flow battery.
Adv. Energy Mater. 1, 511516 (2011).
Herbert, D. & Ulam, J. Electric dry cells and storage batteries. US patent
3,043,896 (1962).
Ji, X. & Nazar, L.F. Advances in Li-S batteries. J.Mater. Chem.
20, 98219826 (2010).
Ji, X., Lee, K.T. & Nazar, L.F. A highly ordered nanostructured carbonsulphur
cathode for lithiumsulphur batteries. Nature Mater. 8, 500506 (2009).
Hassoun, J. & Scrosati, B. A high-performance polymer tin sulfur lithium
ion battery. Angew. Chem. Int. Ed. 49, 23712374 (2010).
Ji, X., Evers, S., Black, R. & Nazar, L.F. Stabilizing lithiumsulphur
cathodes using polysulphide reservoirs. Nature Commun. 2, 325 (2011).

NATURE MATERIALS | VOL 11 | JANUARY 2012 | www.nature.com/naturematerials

2012 Macmillan Publishers Limited. All rights reserved.

27

REVIEW ARTICLE

NATURE MATERIALS DOI: 10.1038/NMAT3191

20. Jeong, S.S. etal. Electrochemical properties of lithium sulfur cells using
PEO polymer electrolytes prepared under three different mixing conditions.
J.Power Sources 174, 745750 (2007).
21. Wang, J.Z. etal. Sulfurgraphene composite for rechargeable lithium batteries.
J.Power Sources 196, 70307034 (2011).
22. Wang, J. etal. Sulfur-mesoporous carbon composites in conjunction with
a novel ionic liquid electrolyte for lithium rechargeable batteries. Carbon
46, 229235 (2008).
23. Peled, E., Sternberg, Y., Gorenshtein, A. & Lavi, Y. Lithiumsulfur
battery: Evaluation of dioxolane-based electrolytes. J.Electrochem. Soc.
136, 16211625 (1989).
24. Aurbach, D. etal. On the surface chemical aspects of very high energy density,
rechargeable Lisulfur batteries. J.Electrochem. Soc. 156, A694-A702 (2009).
25. Abraham, K.M. & Jiang, Z. A polymer electrolyte-based rechargeable lithium/
oxygen battery. J.Electrochem. Soc. 143, 15 (1996).
26. Visco, S.J., Katz, B.D., Nimon, Y.S. & De Jonghe, L.C. Li/air non-aqueous
batteries. US patent 20070117007 (2007).
27. Littauer, E.L. & Tsai, K.C. Anodic behavior of lithium in aqueous-electrolytes.
J.Electrochem. Soc. 123, 771776 (1976).
28. Girishkumar, G., McCloskey, B., Luntz, A.C., Swanson, S. & Wilcke, W.
Lithiumair battery: Promise and challenges. J.Phys. Chem. Lett.
1, 21932203 (2010).
29. Kraytsberg, A. & Ein-Eli, Y. Review on Liair batteriesopportunities,
limitations and perspective. J.Power Sources 196, 886893 (2010).
30. Zhang, J-G. & Bruce, P.G. in Handbook of Batteries (eds Linden, D. &
Reddy, T.B.) 38.4638.73 (McGraw-Hill, 2010).
31. Mikhaylik, Y., Kovalev, I., Xu, J. & Schock, R. Rechargeable LiS
battery with specific energy 350Wh/kg and specific power 3000W/kg.
Meet. Abstr. Electrochem. Soc. 801, 112 (2008).
32. Mikhaylik, Y.V. etal. High energy rechargeable LiS cells for EV application:
Status, remaining problems, and solutions. Meet. Abstr. Electrochem. Soc.
902, 216 (2009).
33. Pistoia, G. Batteries for Portable Devices (Elsevier, 2005).
34. Anderman, M. PHEV and EV Battery Technology Status and Vehicle and
Battery Market Outlook (AABC Europe, 2011).
35. Zhang, S.S., Foster, D. & Read, J. Discharge characteristic of a non-aqueous
electrolyte Li/O2 battery. J.Power Sources 195, 12351240 (2010).
36. Laoire, C.O., Mukerjee, S., Abraham, K.M., Plichta, E.J. &
Hendrickson, M.A. Elucidating the mechanism of oxygen reduction
for lithiumair battery applications. J.Phys. Chem. C
113, 2012720134 (2009).
37. Lu, Y-C., Gasteiger, H.A., Parent, M.C., Chiloyan, V. & Shao-Horn, Y.
The influence of catalysts on discharge and charge voltages of rechargeable
Lioxygen batteries. Electrochem. Solid State 13, A69A72 (2010).
38. Trahey, L. etal. Activated lithium-metal-oxides as catalytic electrodes for
LiO2 cells. Electrochem. Solid State 14, A64A66 (2011).
39. Stevens, P. etal. Development of a lithium air rechargeable battery. ECS Trans.
28, 112 (2010).
40. Hasegawa, S. etal. Study on lithium/air secondary batteries-stability
of NASICON-type lithium ion conducting glassceramics with water.
J.Power Sources 189, 371377 (2009).
41. Zhang, T. etal. Stability of a water-stable lithium metal anode for a
lithiumair battery with acetic acid-water solutions. J.Electrochem. Soc.
157, A214A218 (2010).
42. Laoire, C.O., Mukerjee, S., Abraham, K.M., Plichta, E.J. & Hendrickson, M.A.
Influence of nonaqueous solvents on the electrochemistry of oxygen in the
rechargeable lithiumair battery. J.Phys. Chem. C 114, 91789186 (2010).
43. Laoire, C.O., Mukerjee, S., Plichta, E.J., Hendrickson, M.A. & Abraham, K.M.
Rechargeable lithium/TEGDME-LiPF6/O2 battery. J.Electrochem. Soc.
158, A302A308 (2011).
44. Read, J. Characterization of the lithium/oxygen organic electrolyte battery.
J.Electrochem. Soc. 149, A1190A1195 (2002).
45. Read, J. etal. Oxygen transport properties of organic electrolytes
and performance of lithium/oxygen battery. J.Electrochem. Soc.
150, A1351A1356 (2003).
46. Lu, Y-C. etal. Platinumgold nanoparticles: A highly active bifunctional
electrocatalyst for rechargeable lithiumair batteries. J.Am. Chem. Soc.
132, 1217012171 (2010).
47. Lu, Y-C., Gasteiger, H.A., Crumlin, E., Robert McGuire, J. &
Shao-Horn, Y. Electrocatalytic activity studies of select metal surfaces
and implications in Liair batteries. J.Electrochem. Soc.
157, A1016A1025 (2010).
48. Lu, Y-C., Gasteiger, H.A. & Shao-Horn, Y. Method development to evaluate
the oxygen reduction activity of high-surface-area catalysts for Liair batteries.
Electrochem. Solid State 14, A70A74 (2011).
49. Ogasawara, T., Debart, A., Holzapfel, M., Novak, P. & Bruce, P.G.
Rechargeable Li2O2 electrode for lithium batteries. J.Am. Chem. Soc.
128, 13901393 (2006).
28

50. Dbart, A., Bao, J., Armstrong, G. & Bruce, P.G. An O2 cathode for rechargeable
lithium batteries: The effect of a catalyst. J.Power Sources 174, 11771182 (2007).
51. Dbart, A., Paterson, A., Bao, J. & Bruce, P. -MnO2 nanowires: A catalyst
for the O2 electrode in rechargeable lithium batteries. Angew. Chem. Int. Ed.
47, 45214524 (2008).
52. Kuboki, T., Okuyama, T., Ohsaki, T. & Takami, N. Lithiumair batteries using
hydrophobic room temperature ionic liquid electrolyte. J.Power Sources
146, 766769 (2005).
53. Beattie, S.D., Manolescu, D.M. & Blair, S.L. High-capacity lithiumair
cathodes. J.Electrochem. Soc. 156, A44A47 (2009).
54. Yang, X-H., He, P. & Xia, Y-Y. Preparation of mesocellular carbon foam
and its application for lithium/oxygen battery. Electrochem. Commun.
11, 11271130 (2009).
55. Yang, X-H. & Xia, Y-Y. The effect of oxygen pressures on the electrochemical
profile of lithium/oxygen battery. J.Solid State Electr. 14, 109114 (2010).
56. Zhang, J., Xu, W., Li, X. & Liu, W. Air dehydration membranes for nonaqueous
lithiumair batteries. J.Electrochem. Soc. 157, A940A946 (2010).
57. Zhang, J., Xu, W. & Liu, W. Oxygen-selective immobilized liquid membranes
for operation of lithiumair batteries in ambient air. J.Power Sources
195, 74387444 (2010).
58. Lu, Y.C. etal. The discharge rate capability of rechargeable LiO2 batteries.
Energ. Environ. Sci. 4, 29993007 (2011).
59. Mitchell, R.R., Gallant, B.M., Thompson, C.V. & Shao-Horn, Y. Allcarbon-nanofiber electrodes for high-energy rechargeable Li-O2 batteries.
Energ. Environ. Sci. 4, 29522958 (2011).
60. Xu, W., Xiao, J., Wang, D., Zhang, J. & Zhang, J-G. Crown ethers in nonaqueous
electrolytes for lithium/air batteries. Electrochem. Solid St. 13, A48A51 (2010).
61. Wang, D., Xiao, J., Xu, W. & Zhang, J-G. High capacity pouch-type Liair
batteries. J.Electrochem. Soc. 157, A760A764 (2010).
62. Zhang, J-G., Wang, D., Xu, W., Xiao, J. & Williford, R.E. Ambient operation of
Li/air batteries. J.Power Sources 195, 43324337 (2010).
63. Xiao, J. etal. Optimization of air electrode for Li/air batteries.
J.Electrochem. Soc. 157, A487A492 (2010).
64. Aurbach, D., Daroux, M., Faguy, P. & Yeager, E. The electrochemistry of
noble metal electrodes in aprotic organic solvents containing lithium salts.
J.Electroanal. Chem. 297, 225244 (1991).
65. Mizuno, F., Nakanishi, S., Kotani, Y., Yokoishi, S. & Iba, H. Rechargeable
Liair batteries with carbonate-based liquid electrolytes. Electrochemistry
78, 403405 (2010).
66. Xu, W. etal. Investigation on the charging process of Li2O2-based
air electrodes in LiO2 batteries with organic carbonate electrolytes.
J.Power Sources 196, 38943899 (2011).
67. Freunberger, S.A. etal. Fundamental mechanism of the lithiumair battery.
Meet. Abstr. - Electrochem. Soc. 1003, 399 (2010).
68. Veith, G.M., Dudney, N.J., Howe, J. & Nanda, J. Spectroscopic
characterization of solid discharge products in Li-air cells with aprotic
carbonate electrolytes. J.Phys. Chem. C 115, 1432514333 (2011).
69. Freunberger, S.A. etal. Reactions in the rechargeable lithiumO2 battery with
alkyl carbonate electrolytes. J.Am. Chem. Soc. 133, 80408047 (2011).
70. Freunberger, S.A. etal. The lithiumoxygen battery with ether-based
electrolytes. Angew. Chem. Int. Ed. 50, 86098613 (2011).
71. McCloskey, B.D., Bethune, D.S., Shelby, R.M., Girishkumar, G. &
Luntz, A.C. Solvents critical role in nonaqueous lithiumoxygen battery
electrochemistry. J.Phys. Chem. Lett. 2, 11611166 (2011).
72. Hassoun, J., Croce, F., Armand, M. & Scrosati, B. Investigation of
the O2 electrochemistry in a polymer electrolyte solid-state cell.
Angew. Chem. Int. Ed. 50, 29993002 (2011).
73. Peng, Z. etal. Oxygen reactions in a non-aqueous Li+ electrolyte.
Angew. Chem. Int. Ed. 50, 63516355 (2011).
74. Bard, F., Bruce, P.G., Freunberger, S.A. & Hardwick, L.J. Cathode
catalyst for rechargeable metalair & rechargeable metalair battery.
JPO patent 059494 (2010).
75. Bard, F., Bruce, P.G., Freunberger, S.A., Chen, Y. & Hardwick, L.J.
Catalyst loaded onto carbon for rechargeable nonaqueous metalair battery.
JPO patent 053888 (2011).
76. Cheng, H. & Scott, K. Carbon-supported manganese oxide nanocatalysts for
rechargeable lithiumair batteries. J.Power Sources 195, 13701374 (2010).
77. Giordani, V., Freunberger, S.A., Bruce, P.G., Tarascon, J-M. & Larcher, D.
H2O2 decomposition reaction as selecting tool for catalysts in LiO2 cells.
Electrochem. Solid St. 13, A180A183 (2010).
78. Sawyer, D.T. & Roberts, J.L. Electrochemistry of oxygen and superoxide
ion in dimethylsulfoxide at platinum, gold and mercury electrodes.
J.Electroanal. Chem. 12, 90101 (1966).
79. Kumar, B. etal. A solid-state, rechargeable, long cycle life lithium-air battery.
J.Electrochem. Soc. 157, A50A54 (2010).
80. Wang, Y. & Zhou, H. A lithiumair battery with a potential to
continuously reduce O2 from air for delivering energy. J.Power Sources
195, 358361 (2010).
NATURE MATERIALS | VOL 11 | JANUARY 2012 | www.nature.com/naturematerials

2012 Macmillan Publishers Limited. All rights reserved.

REVIEW ARTICLE

NATURE MATERIALS DOI: 10.1038/NMAT3191


81. He, P., Wang, Y. & Zhou, H. A Li-air fuel cell with recycle aqueous electrolyte
for improved stability. Electrochem. Commun. 12, 16861689 (2010).
82. He, P., Wang, Y.G. & Zhou, H.S. The effect of alkalinity and temperature
on the performance of lithiumair fuel cell with hybrid electrolytes.
J.Power Sources 196, 56115616 (2011).
83. Wang, Y.G. & Zhou, H.S. A lithiumair fuel cell using copper to catalyze
oxygen-reduction based on copper-corrosion mechanism. Chem. Commun.
46, 63056307 (2010).
84. Suntivich, J. etal. Design principles for oxygen-reduction activity on
perovskite oxide catalysts for fuel cells and metalair batteries. Nature Chem.
3, 546550 (2011).
85. Cheon, S-E. etal. Rechargeable lithium sulfur battery. J.Electrochem. Soc.
150, A800A805 (2003).
86. Choi, Y-J., Kim, K-W., Ahn, H-J. & Ahn, J-H. Improvement of cycle property of
sulfur electrode for lithium/sulfur battery. J.Alloy Compd. 449, 313316 (2008).
87. Marston, J.M. & Brummer, S.B. Formation of lithium polysulfides in aprotic
media. J. Inorg. Nucl. Chem. 39, 17611766 (1977).
88. Yamin, H. & Peled, E. Electrochemistry of a nonaqueous lithium/sulfur cell.
J.Power Sources 9, 281287 (1983).
89. Ryu, H.S., Guo, Z., Ahn, H.J., Cho, G.B. & Liu, H. Investigation of
discharge reaction mechanism of lithium liquid electrolyte sulfur battery.
J.Power Sources 189, 11791183 (2009).
90. Yamin, H., Gorenshtein, A., Penciner, J., Sternberg, Y. & Peled, E. Lithium
sulfur battery oxidation reduction-mechanisms of polysulphides in THF
solutions. J.Electrochem. Soc. 135, 10451048 (1988).
91. Mikhaylik, Y.V. & Akridge, J.R. Polysulfide shuttle study in the Li/S battery
system. J.Electrochem. Soc. 151, A1969A1976 (2004).
92. Degott, P., Polymere Carbone-Soufre Synthese et Proprietes Electrochimiques
PhD Thesis, lInstitut National Polytechnique de Grenoble (1986).
93. Visco, S.J., Mailhe, C.C., Jonghe, L.C.D. & Armand, M.B. A novel
class of organosulfur electrodes for energy storage. J.Electrochem. Soc.
136, 661664 (1989).
94. Liu, M., Visco, S.J. & Jonghe, L.C.D. Electrochemical properties of organic
disulfide/thiolate redox couples. J.Electrochem. Soc. 136, 25702575 (1989).
95. Kiya, Y., Iwata, A., Sarukawa, T., Henderson, J.C. & Abrua, H.D. Poly[dithio2,5-(1,3,4-thiadiazole)] (PDMcT)-poly(3,4-ethylenedioxythiophene)
(PEDOT) composite cathode for high-energy lithium/lithium-ion
rechargeable batteries. J.Power Sources 173, 522530 (2007).
96. Kiya, Y., Henderson, J.C., Hutchison, G.R. & Abruna, H.D. Synthesis,
computational and electrochemical characterization of a family of functionalized
dimercaptothiophenes for potential use as high-energy cathode materials for
lithium/lithium-ion batteries. J.Mater. Chem. 17, 43664376 (2007).
97. Xu, G.X., Bi, L.Q., Yu, T. & Wen, L. PVC disulfide as cathode materials for
secondary lithium batteries. Chinese J.Polym. Sci. 24, 307313 (2006).
98. Rauh, R.D., Abraham, K.M., Pearson, G.F., Surprenant, J.K. &
Brummer, S.B. A lithium/dissolved sulfur battery with an organic electrolyte.
J.Electrochem. Soc. 126, 523527 (1979).
99. Yamin, H., Penciner, J., Gorenshtain, A., Elam, M. & Peled, E. The
electrochemical behavior of polysulfides in tetrahydrofuran. J.Power Sources
14, 129134 (1985).
100. Peled, E., Gorenshtein, A., Segal, M. & Sternberg, Y. Rechargeable lithium
sulfur battery. J.Power Sources 26, 269271 (1989).
101. Tobishima, S-I., Yamamoto, H. & Matsuda, M. Study on the reduction
species of sulfur by alkali metals in nonaqueous solvents. Electrochim. Acta
42, 10191029 (1997).
102. Chu, M-Y. Liquid electrolyte lithiumsulfur batteries. US patent 6030720 (2000).
103. Shin, J.H. & Cairns, E.J. Characterization of N-methyl-N-butylpyrrolidinium
bis(trifluoromethanesulfonyl)imide-LiTFSI-tetra(ethylene glycol)
dimethyl ether mixtures as a Li metal cell electrolyte. J.Electrochem. Soc.
155, A368A373 (2008).
104. Choi, J-W. etal. Rechargeable lithium/sulfur battery with suitable mixed liquid
electrolytes. Electrochim. Acta 52, 20752082 (2007).
105. Marmorstein, D. etal. Electrochemical performance of lithium/sulfur cells
with three different polymer electrolytes. J.Power Sources 89, 219226 (2000).
106. Wang, J.L., Yang, J., Xie, J.Y., Xu, N.X. & Li, Y. Sulfurcarbon nanocomposite as cathode for rechargeable lithium battery based on gel electrolyte.
Electrochem. Commun. 4, 499502 (2002).
107. Hayashi, A., Ohtomo, T., Mizuno, F., Tadanaga, K. & Tatsumisago, M.
All-solid-state Li/S batteries with highly conductive glassceramic electrolytes.
Electrochem. Commun. 5, 701705 (2003).
108. Yang, Y. etal. New nanostructured Li2S/silicon rechargeable battery with high
specific energy. Nano Lett. 10, 14861491 (2010).
109. Han, S-C. etal. Effect of multiwalled carbon nanotubes on electrochemical
properties of lithium/sulfur rechargeable batteries. J.Electrochem. Soc.
150, A889A893 (2003).
110. Zheng, W., Liu, Y.W., Hu, X.G. & Zhang, C.F. Novel nanosized adsorbing
sulfur composite cathode materials for the advanced secondary lithium
batteries. Electrochim. Acta 51, 13301335 (2006).

111. Niu, J.J., Wang, J.N., Jiang, Y., Su, L.F. & Ma, J. An approach to
carbon nanotubes with high surface area and large pore volume.
Micropor. Mesopor. Mater. 100, 15 (2007).
112. Yuan, L., Yuan, H., Qiu, X., Chen, L. & Zhu, W. Improvement of cycle
property of sulfur-coated multi-walled carbon nanotubes composite cathode
for lithium/sulfur batteries. J.Power Sources 189, 11411146 (2009).
113. Song, M-S. etal. Effects of nanosized adsorbing material on electrochemical
properties of sulfur cathodes for Li/S secondary batteries. J.Electrochem. Soc.
151, A791A795 (2004).
114. Choi, Y.J. etal. Electrochemical properties of sulfur electrode containing
nano Al2O3 for lithium/sulfur cell. Phys. Scripta T129, 6265 (2007).
115. Wang, J., Yang, J., Xie, J. & Xu, N. A novel conductive polymersulfur
composite cathode material for rechargeable lithium batteries. Adv. Mater.
14, 963965 (2002).
116. Yu, X-g. etal. Lithium storage in conductive sulfur-containing polymers.
J.Electroanal. Chem. 573, 121128 (2004).
117. Wang, J. etal. Sulphur-polypyrrole composite positive electrode materials for
rechargeable lithium batteries. Electrochim. Acta 51, 46344638 (2006).
118. Lai, C., Gao, X.P., Zhang, B., Yan, T.Y. & Zhou, Z. Synthesis and
electrochemical performance of sulfur/highly porous carbon composites.
J.Phys. Chem. C 113, 47124716 (2009).
119. Liang, C., Dudney, N.J. & Howe, J.Y. Hierarchically structured sulfur/carbon
nanocomposite material for high-energy lithium battery. Chem. Mater.
21, 47244730 (2009).
120. Jayaprakash, N., Shen, J., Moganty, S.S., Corona, A. & Archer, L.A. Porous
hollow carbon@sulfur composites for high-power lithiumsulfur batteries.
Angew. Chem. Int. Ed. 50, 59045908 (2011).
121. Li, S., Xie, M., Liu, J., Wang, H. & Yan, H. Layer structured sulfur/expanded
graphite composite as cathode for lithium battery. Electrochem. Solid St.
14, A105A107 (2011).
122. Cao, Y. etal. Sandwich-type functionalized graphene sheetsulfur
nanocomposite for rechargeable lithium batteries. Phys. Chem. Chem. Phys.
13, 76607665 (2011).
123. Wu, F. etal. Sulfur/polythiophene with a core/shell structure: Synthesis and
electrochemical properties of the cathode for rechargeable lithium batteries.
J.Phys. Chem. C 115, 60576063 (2011).
124. Qiu, L., Zhang, S., Zhang, L., Sun, M. & Wang, W. Preparation and enhanced
electrochemical properties of nano-sulfur/poly(pyrrole-co-aniline) cathode
material for lithium/sulfur batteries. Electrochim. Acta 55, 46324636 (2010).
125. Demir-Cakan, R. etal. Cathode composites for LiS batteries via
the use of oxygenated porous architectures. J.Am. Chem. Soc.
133, 1615416160 (2011).
126. Mirzaeian, M. & Hall, P.J. Preparation of controlled porosity carbon aerogels
for energy storage in rechargeable lithium oxygen batteries. Electrochim. Acta
54, 74447451 (2009).
127. Zhang, G.Q. etal. Lithiumair batteries using SWNT/CNF buckypapers as air
electrodes. J.Electrochem. Soc. 157, A953A956 (2010).
128. Albertus, P. etal. Identifying capacity limitations in the Li/oxygen
battery using experiments and modeling. J.Electrochem. Soc.
158, A343A351 (2011).
129. Mikhaylic, Y.V. Electrolytes for lithium sulfur cells. US patent 7354680 (2008).
130. http://oharacorp.com/pdf/LIC-GC.pdf
131. Imanishi, N. etal. Lithium anode for lithiumair secondary batteries.
J.Power Sources 185, 13921397 (2008).
132. Zhang, T. etal. A novel high energy density rechargeable lithium/air battery.
Chem. Commun. 46, 16611663 (2010).
133. Zhang, T., Imanishi, N., Hirano, A., Takeda, Y. & Yamamoto, O. Stability
of Li/polymer electrolyte-ionic liquid composite/lithium conducting
glass ceramics in an aqueous electrolyte. Electrochem. Solid State
14, A45A48 (2011).
134. Debart, A., Dupont, L., Patrice, R. & Tarascon, J-M. Reactivity of transition
metal (Co, Ni, Cu) sulphides versus lithium: The intriguing case of the copper
sulphide. Solid State Sci. 8, 640651 (2006).
135. Zhang, S.S., Foster, D. & Read, J. A high energy density lithium/sulfur-oxygen
hybrid battery. J.Power Sources 195, 36843688 (2010).
136. http://www.nissanusa.com/leaf-electric-car/specs-features/index#/leafelectric-car/specs-features/index.
137. US Advanced Battery Consortium USABC Goals for Advanced Batteries for EVs
(2006). Available at: http://uscar.org/commands/files_download.php?files_id=27.

Acknowledgements

P.G.B. is indebted to the EPRSC and Toyota Motor Europe for support. The authors wish
to express their thanks to S.Visco, M.Armand and R.Demir-Cakan and the ALISTOREERI members for helpful discussions. P.G.B. and J.M.T. are members of ALISTOREERIEuropean Network of Excellence on Lithium Batteries.

Additional information

The authors declare no competing financial interests.

NATURE MATERIALS | VOL 11 | JANUARY 2012 | www.nature.com/naturematerials

2012 Macmillan Publishers Limited. All rights reserved.

29

ARTICLES

NATURE MATERIALS DOI: 10.1038/NMAT3237

ERRATUM

LiO2 and LiS batteries with high energy storage


Peter G. Bruce, Stefan A. Freunberger, Laurence J. Hardwick and Jean-Marie Tarascon
Nature Materials 11, 1929 (2012); published online: 15 December 2011; corrected after print: 4 January 2012.
In the version of this Review originally published, in Table 1, the values in rows 25 of the Cell voltage column appeared incorrectly;
the full column should have read 3.8, 1.65, 2.2, 3.0 and 3.2. This has now been corrected in the HTML and PDF versions.

172

NATURE MATERIALS | VOL 11 | FEBRUARY 2012 | www.nature.com/naturematerials

2012 Macmillan Publishers Limited. All rights reserved.

Vous aimerez peut-être aussi