Sie sind auf Seite 1von 15

REVIEW ARTICLE

PUBLISHED: 24 JULY 2017 | VOLUME: 2 | ARTICLE NUMBER: 17108

A materials perspective on Li-ion batteries at


extreme temperatures
Marco-Tulio F. Rodrigues, Ganguli Babu, Hemtej Gullapalli, Kaushik Kalaga, Farheen N. Sayed,
Keiko Kato, Jarin Joyner and Pulickel M. Ajayan*

With the continuous upsurge in demand for energy storage, batteries are increasingly required to operate under extreme envi -
ronmental conditions. Although they are at the technological forefront, Li-ion batteries have long been limited to room tem-
perature, as internal phenomena during their operation cause thermal fluctuations. This has been the reason for many battery
explosions in recent consumer products. While traditional efforts to address these issues focused on thermal management strat-
egies, the performance and safety of Li-ion batteries at both low (<20°C) and high (>60°C) temperatures are inherently related
to their respective components, such as electrode and electrolyte materials and the so-called solid-electrolyte interphases. This
Review examines recent research that considers thermal tolerance of Li-ion batteries from a materials perspective, spanning a
wide temperature spectrum (–60 °C to 150 °C). The structural stability of promising cathodes, issues with anode passivation,
and the competency of various electrolyte, binder and current collectors are compared for their thermal workability. The pos-
sibilities offered by each of these cell components could extend the environmental frontiers of commercial Li-ion batteries.

E
nergy storage forms the foundation for success of numerous is impractical. Many applications requiring extreme temperature
commercial products. Though many battery chemistries exist, windows rely on primary lithium thionyl chloride (Li–SOCl2) bat-
Li-ion batteries (LIBs) are at the forefront for rechargeable appli- teries, usable from –60 °C to 150 °C (ref. 5). Despite this impres-
cations, as the combination of high energy density, light weight, and sive thermal resilience, peak performance is only achieved between
low self-discharge makes them highly attractive for portable con- 20 °C and 55 °C, and cell usage is confined to low power conditions.
sumer electronics1. While a large spectrum of consumer applications A LIB with proper formulation can potentially meet and surpass
operate at room temperature, demand for batteries to survive and the temperature window offered by Li–SOCl2 batteries, while also
operate under thermal extremes is rising. Military-grade batteries offering rechargeability, higher power capability and reduced toxic-
are expected to operate from –40 °C to 60 °C, and such LIBs are yet ity. Besides serving diversified applications, expanding the tempera-
to be fully optimized and developed. Electric vehicles require battery ture limits of LIBs would also have a rather beneficial side-effect:
systems capable of having stable performance in both colder regions improvement of consumer devices’ resistance to thermal abuse.
and hot desert conditions. Apart from extending the operability of In designing a battery configuration for a targeted application, it
conventional batteries, task-specific applications also call for energy is critical to understand the workability of the materials being used.
storage at further extremes. Areas like subsurface exploration (such The charge storage mechanism in LIBs involves a synergy of trans-
as for oil and gas), thermal reactors, defensive arsenals and space formations among its components, and temperature plays a deci-
vehicles call for operating temperatures over a wide range. On the sive role. Many recent efforts have considered thermal effects on the
other hand, equipment for cold climates and high-altitude drones individual materials used for LIBs in a piecewise approach. Though
can experience temperatures as low as –60 °C. Batteries might also be it is not guaranteed that a full-cell encompassing each of these frag-
required to operate under a predefined temperature regime. Medical ments would function as desired, it is nevertheless important to
devices, for example, work at room temperature, but are subjected to individually understand the limits of these materials.
120 °C during sterilization. In uses such as energy harvesting, drill- This Review examines recent reports on thermal characteristics
ing sensors and drones, batteries experience dissimilar temperatures of battery components and attempts to present a materials perspec-
during charge and discharge. Besides external influences, internal tive, both at low and high temperature extremes. Reports pertaining
phenomena while operating LIBs also cause thermal fluctuations2, to electrochemical behaviour of electrode materials, development
sometimes unintentionally swinging to hazardous extremes3. Recent of unconventional electrolytes and the role of dormant support
battery explosions in cases such as the Boeing Dreamliner and components (such as binders and current collectors) in the perfor-
Samsung Note7 have raised already existing safety concerns4. mance at extreme temperatures are discussed in subsequent sec-
LIBs are in a unique position to cater to all these applications: tions. Although optimal cell-level performance under demanding
while other cell chemistries have long-lived issues limiting their conditions has yet to be fully achieved, the intrinsic properties of
operability, the range of electrode and electrolyte materials compat- these materials cast favourable prospects over the evolution of this
ible with LIBs can potentially render these devices active across a battery chemistry.
wide temperature range. Among other cell concepts, water-based
technologies, as lead–acid and nickel–metal hydride, are intrinsi- Cathodes
cally limited by the electrolyte to operate between –50 °C and 50 °C The cathode is the transit centre for both Li + and electrons, and
(ref. 5). Extremely high temperatures are compatible with — and these materials are susceptible to undesirable phase transitions.
required by — molten salt batteries, while operation below 90 °C Further, parasitic reactions at the electrode–electrolyte interface in

Department of Material Science and Nanoengineering, Rice University, Houston, Texas 77005, USA. *e-mail: ajayan@rice.edu

NATURE ENERGY 2, 17108 (2017) | DOI: 10.1038/nenergy.2017.108 | www.nature.com/natureenergy 1


©
2
0
1
7
M
a
c
im
lla
n
P
u
b
ils
h
re
s
L
im
it
e
d
,
p
a
rt
o
fp
S
ri
n
g
re
N
a
t
u
r
e
.
A
ll
ri
g
h
ts
r
e
s
e
r
v
e
d
.
REVIEW ARTICLE NATURE ENERGY

a b c d
100
80

retention (%)
Capacity
LFP
60 LCO NMC
LMO
40 Ambient
Ambient (ref. 8) Ambient (ref. 10) Ambient (ref. 32)
20 (our data)
60 ºC (ref. 8) 60 ºC (ref. 10) 55 ºC (ref. 32) 55 ºC (ref. 21)
0
0 25 50 0 25 50 0 25 50 0 25 50
Cycle number (#)

e f Cation 2º phase
Lattice
mixing formation
strain LCO
(Li+–Ni2+)
Metal NMC
dissolution Metal
Oxygen Li+
Li+ dissolution
evolution
Co Slow charge transfer Mn/Ni/Co
Slow charge transfer
and Li+ di¦usivity and Li+ di¦usivity O2–
O2–

g LMO h High
Mn3+ LFP
Mn4+ internal
Li+ impedance
2º phase
formation Mn Li+ Poor e–
O 2- Fe conductivity
Jahn–Teller
distortion P4+ Sluggish
Co-existence of Metal O2– Li transport
two phases at HV dissolution

i j 4.4
260
4.2
240
4.0
220 25 °C
3.8
Temperature (°C)

E vs (Li/Li+)/V

200 Li1–xCoO2 3.6


–20 °C
Li1–xNiO2
180 3.4 4
Li1–xNi0.8Co0.2O2
3.2 1 Pristine
160 Li1–xMn 2O4 2 D500
3.0 3 D1100
Li1–xFePO4 4 D2200 2
140 Mn-rich NMCs 4
2.8 5 D3500 5
Mn-deficient NMCs 1 3
120 2.6 123 5
0.0 0.2 0.4 0.6 0.8 1.0 –20 0 20 40 60 80 100 120 140 160 180 200
Fraction of Li in cathode materials Capacity/mAh g–1

Figure 1 | Temperature-dependent performance constraints in cathode materials. a–h, Cycling plots (a–d) show ambient and elevated temperature
performance while schematics (e–h) show structural instabilities that get aggravated at high (text in red) and low (text in blue) temperature for
representative cathode materials. Layered oxide framework cathodes LiCoO 2 (LCO) (a,e); LiNi1/3Mn1/3Co1/3O2 (NMC) (b,f); spinel LiMn2O4 (c,g); olivine
phosphate LiFePO4 (d,h). The data points for plots a–d have been deduced from previous reports as indicated in the panels. i, Thermal stability as a
function of lithium content in cathode materials. The temperature values indicate the onset of reaction between carbonate-based electrolyte and the
electrode material at the respective state-of-charge. The connected data points are derived from DSC curves, while shaded regions compare influence of
the Mn concentration on the stability of NMC cathode11,110. j, Performance of LiNi0.5Co0.2Mn0.3O2 (here denoted as NCM523) at low and room temperatures.
Substitution of an optimum concentration of Al inhibits impedance rise and accelerates lithium ion diffusion at low temperature, which translates into
improved performance. Panel j reproduced with permission from ref. 26 (Elsevier).

deep-charged conditions (delithiated state) are likely to destabilize the self-accelerating reactivity of decayed electrodes with flam-
the cathode’s structure. mable electrolyte at elevated temperatures poses safety concerns6
The reported cyclic stability of well-known cathode materi- (Fig. 1i). Contrarily, sluggish kinetics limits cathode performance
als at high temperature and their source of fragility at temperature below 0 °C (Fig. 1j). Although the development of high-energy-
extremes is schematically shown in Figs 1a–h. The listed issues are density cathodes is critical for many applications, growing demand
minimal or negligible at room temperature but accelerated upon cell for LIBs to survive/operate under extreme conditions has fostered
exposure to high temperatures, increasing the extent of metal dis- considerable research interest in probing the characteristics of cath-
solution and oxygen evolution. These processes are generally irre- ode materials with temperature as a variable. Table 1 summarizes
versible, leading to performance degradation owing to loss of active representative reports pertaining to known cathode materials and
mass and instability in the structural composition. Additionally, their modifications.

2 NATURE ENERGY 2, 17108 (2017) | DOI: 10.1038/nenergy.2017.108 | www.nature.com/natureenergy


©
2
0
1
7
M
a
c
im
lla
n
P
u
b
ils
h
re
s
L
im
it
e
d
,
p
a
rt
o
fp
S
ri
n
g
re
N
a
t
u
r
e
.
A
ll
ri
g
h
ts
r
e
s
e
r
v
e
d
.
NATURE ENERGY REVIEW ARTICLE
Table 1 | Electrochemical performance of cathode materials at extreme temperatures.
Cathodes Temperature (°C) Capacity Number of cycles Capacity Electrolyte and anode
(mAhg–1) retention (%)
Layered oxide cathodes
AlPO4/LiCoO2 (ref. 7) 60 180 30 79 Organic and lithium
LiNi0.33Co0.33Mn0.33O2 (ref. 9) 60 175 100 95.3 Organic and Li4Ti5O12
Al doping/ LiNi 0.5Co0.2Mn0.3O2 –20 106 - - Organic and lithium
(ref. 26)

Spinel oxide cathodes


LTO-coated LiMn2O4 (ref. 94) 60 132 100 97 Organic and lithium
LiMn1.5Ni0.5O3.8F0.2 (ref. 18) 60 110 50 40 Organic and lithium
Li1+xMn2–xO4 (ref. 17) 120 80 20 50 RTIL and Li4Ti5O12

Phosphate cathodes
C-LiFePO4 (ref. 29) 60 160 40 93 Organic and lithium
LiFePO4 (ref. 95) 120 160 20 94 PC-LiBOB and lithium
LiFePO4 (ref. 20) 250 - 8 - Molten LiTFSI and lithium
Nano-LiFePO4/C (ref. 96) –20 120 100 99 Organic and lithium
LiFePO4/C (ref. 97) –40 54.7 - 39 Organic and lithium

Layered oxide framework cathodes. Lithium cobalt oxide (LCO), justifying its adoption in the Panasonic batteries powering the latest
the commercially established cathode, belongs to the two-dimen- Tesla vehicles13. However, the energetic nature14 of NCA conditions
sional layered materials family (LiMO 2, M = Co,Ni,Mn), wherein its use in high-capacity cells to strict state-of-health monitoring and
Li+ occupy the sites between two O–M–O layers. Though it shows safety control.
satisfactory performance at room temperature with intercalation/ Another approach to improve high-temperature stability is to
de-intercalation of 0.5 Li per unit formula, structural instability minimize the electrode–electrolyte contact through surface engi-
in the deeply-charged state and fragile Co–O bonding limits its neering. Hydrofluoric acid (HF) originating from electrolyte decay
use to 60 °C (refs 7,8) (Fig. 1a). The intermediate oxidation state favours metal dissolution, contributing to oxygen evolution from
of cobalt (+3.5) in charged Li 1–xCoO2 makes the crystal suscep- the cathode. Coating with electrochemically inert materials, such as
tible to change from the native hexagonal phase to a monoclinic oxides and phosphates, has proven effective as the bulk of the cath-
structure, releasing oxygen due to metal dissolution in the electro- ode is shielded from interfacial degradations 6. Although structural
lyte, in a process that is aggravated beyond 60 °C. Other layered modifications have been successful in expanding the temperature
oxides present similar stability limitations, as the structural breath- range of layered oxides by a marginal extent, their intrinsic fragility
ing during Li-ion intercalation/de-intercalation generates residual makes them vulnerable to thermal degradation.
stresses in the crystallographic z-direction. Temperature-provoked
cation exchange and structural transformation to spinel structures Spinel-structured cathodes. Unlike in layered materials, changes
under deeply-charged conditions have written off the commer- in unit cell lattice parameters during charge–discharge are iso-
cial use of materials such as LiNiO2 and LiMnO2 at the early stage tropic in spinel LiMn2O4 (LMO). This material exhibits good ther-
of development. mal stability and rate capability, owing to its strong edge-shared
In the case of mixed metal oxides, there is a positive synergy (Mn2)O4 octahedral lattice and three-dimensional Li-ion conduc-
between their constituting elements. LiNi1/3Mn1/3Co1/3O2 (NMC), tion pathways. Thermogravimetric analysis of deeply charged LMO
for instance, exhibits high capacity (Ni 2+), minimal cationic dis- (λ-MnO2) showed that oxygen release usually occurs at tempera-
order (Co3+) and extended thermal stability (Mn 4+) (refs 9,10). tures higher than for layered materials, indicating the robustness of
Consequently, NMC has demonstrated good capacity retention at spinels. Although λ-MnO2 does not evolve oxygen until ~400 °C,
60 °C, presenting high coulombic efficiencies for over 100 cycles9. it converts into inactive MnO2 phases at as low as 190 °C (ref. 14).
These features motivated NMC’s incursion in the commercial mar- Electrolyte oxidation at 4.0 V3+and metal dissolution due to dispro-
ket, with growing use in cells for portable electronics. At elevated portionation reactions (2Mn → Mn4+ + Mn2+ ) are
(solid) (solid) (liquid)
temperature, challenges for lasting cyclability are metal dissolution serious concerns for LMO. Moreover, Jahn−Teller distortion due
into electrolyte, and minimization of cation mixing, as Li+ and Ni2+ to the presence of excess Mn 3+ at 3.0 V leads to phase transition
exhibit similar ionic radii (Fig. 1f). from cubic to tetragonal (Fig. 1g). As these processes become more
The presence of electrochemically inactive species may exhibit prominent at higher temperatures, structural and surface engineer-
a pillaring effect, stabilizing the active cations at their respective ing are required to stabilize the cathode.
crystallographic positions: high manganese concentrations in the Like layered materials, structural stabilization can be attained
lattice, for instance, stabilize the crystal structure of delithiated using coatings and doping 6,15. Surface protection of LMO using
NMC (Fig. 1i; ref. 11). Similarly, LiNi0.8Co0.15Al0.05O2 (NCA) deliv- inert oxides, fluorides and phosphates is an effective approach to
ers high stable capacity (~200 mAhg–1), as doping with inert Al improve performance up to 60 °C, as these materials can act as
enhances charge retention, facilitating maximum utilization of the HF scavengers. However, insulating coatings inflict structural mis-
active transition metal. However, capacity fade has been observed match, creating stacking faults that can obstruct Li+ diffusion16. The
for NCA at 40 °C to 70 °C, due to abrupt surface film growth and surface layer thickness should be in the 1–2 nm range, effectively
micro-cracks at grain boundaries12. NCA is highly cost-effective, blocking solvent molecules while allowing Li-ion conduction.

NATURE ENERGY 2, 17108 (2017) | DOI: 10.1038/nenergy.2017.108 | www.nature.com/natureenergy 3


©
2
0
1
7
M
a
c
im
lla
n
P
u
b
ils
h
re
s
L
im
it
e
d
,
p
a
rt
o
fp
S
ri
n
g
re
N
a
t
u
r
e
.
A
ll
ri
g
h
ts
r
e
s
e
r
v
e
d
.
REVIEW ARTICLE NATURE ENERGY

Li+
E"ects at high temperature Solvent E"ects at low temperature
a c Metal ion e g

–9

LgD (cm2s–1)
SEI
Li dendrite
Reaction with growth –10
metal cations
Delithiated
Lithiated

–11
3 3.4 3.8 4.2
1000/T (k–1)
b d f h
Increase in
charge-transfer 21 °C 5 °C –10 °C –23 °C
and SEI
0.00

(V versus Li/Li+)
Anode polarization
Gas evolution resistances

omposition Ω
change –0.08 Anode active
SEI material
thickening Amorphous carbon
–0.12
Graphite
0.0034 0.0036 0.0038 0.0040
1/T(K–1)

Figure 2 | Challenges imposed by temperature on the performance of graphite anodes. a–f, The many interfacial processes that might take place at
the electrode. Elevated temperatures can damage the SEI (a), promoting further decomposition of the electrolyte to form a thicker and more resistive
layer (b). The passivation film can also be negatively affected by reactive species generated on the cathode (c), which contributes to permanent changes
in SEI composition (d). At the low temperature extreme, lithium plating becomes very problematic (e), and reaction kinetics are negatively affected
by the increased Rct and the SEI’s blocking character. g, Temperature dependence of the apparent chemical diffusion coefficient (D) of Li+ in graphite.
There is a large decrease in D below –20 °C, when the solid-state diffusion also becomes much slower during the lithiation process, with deep effects
on rate capability. h, Anodic polarization for Li+ insertion at different temperatures. The cell overpotential rises quickly at low temperatures, leading to
deposition of metallic lithium at the electrode surface (inset). The low polarization observed for amorphous carbon renders this material much safer at
low temperatures. Panels reproduced with permission from: g, ref. 34 (Elsevier); h ref. 29 (Elsevier).

Doping Mn-sites with suitable ions stabilizes the lattice by expand- involving elevated temperatures, because of features such as: high
ing the regular MO6 octahedra, improving accommodation of thermal and structural stability, as decomposition occurs above
Mn3+ → Mn4+ transitions, even at 55 °C (ref. 15). For this, several 400 °C; oxygen release unlikely from PO43–; and minimal volume
isovalent (Cr3+, Al3+) and aliovalent (Mg2+, Ti4+) dopants have been mismatch between LiFePO4/FePO4 phases19,20, improving struc-
successfully employed. Use of alternative electrolytes can also be tural integrity after complete Li removal. The resilience of LFP in a
beneficial for applications in more demanding environments, cell environment has been highlighted by its use in an LIB at 250 °C
as cycling at 120 °C has been demonstrated using ionic liquids 17. with molten salt electrolyte20.
Despite its relative thermal stability and historical prominence, Despite these favourable properties, LFP’s poor ionic and elec-
the relatively low capacity (100–120 mAhg–1) of LMO bars it from tronic conductivities still preclude utilization of micron-sized active
widespread commercial use. material particles, especially without conductive coating (Fig. 1h).
Recently, the high-voltage spinel LiMn1.5Ni0.5O4 has been in In addition, LFP is natively susceptible to metal dissolution, requir-
the spotlight due to its high energy density. However, introduc- ing conformal coating with carbon, metal oxides or polymers 19,21
tion of Ni into the lattice reduces the thermal stability, provok- to achieve considerable cycle life. Though the low energy density
ing oxygen release at temperatures much lower than for LMO 18. diminishes the significance of LFP for the consumer electronic
Further, structural decomposition is likely to be more prominent market, its higher power density and long cycle and calendar lives
at high temperature during the complete removal of lithium at make this material rather attractive for specific industrial and sta-
high voltages. Although more development is required to confer tionary applications. LFP has also been employed by Bolloré Group
thermal stability to high-voltage chemistries, spinels are expected in solid-state batteries employing polymer electrolyte and lithium
to be comparatively more thermally resilient than layered oxides, metal anodes for electric vehicles.
with the major hurdles being metal dissolution and Jahn–Teller Recently, other phosphates, such as VOPO 4 and its lithiated
distortion. compounds (LiVOPO4, Li2VOPO4), have been found to exhibit
high thermal stability14. Like LFP, VOPO4 does not evolve O2 upon
Phosphate cathodes. The limited structural stability of layered heating, and only reacts with electrolyte beyond 200 °C, making it
and spinel cathodes arises mostly from the fragility of M–O bonds inherently safe. Li3V2(PO4)3 (LVP) is another candidate for applica-
constructing the lattice. Polyanion-based phosphates provide more tions involving a wide range of temperatures, asserting good Li-ion
resilient frameworks, since oxygen atoms are covalently bonded to kinetics. However, the high operating voltages might render LVP
phosphorous instead of transition metals, facilitating charge distri- susceptible to metal dissolution at elevated temperatures, and dop-
bution19. Of the known cathode structures, phosphates in general ing might play an essential role in stabilizing the charge density over
and LiFePO4 (LFP) specifically are leading choices for applications vanadium atoms.

4 NATURE ENERGY 2, 17108 (2017) | DOI: 10.1038/nenergy.2017.108 | www.nature.com/natureenergy


©
2
0
1
7
M
a
c
im
lla
n
P
u
b
ils
h
re
s
L
im
it
e
d
,
p
a
rt
o
fp
S
ri
n
g
re
N
a
t
u
r
e
.
A
ll
ri
g
h
ts
r
e
s
e
r
v
e
d
.
NATURE ENERGY REVIEW ARTICLE
Table 2 | Electrolyte components to extend the temperature range of Li-ion batteries.
Low temperature (>–60°C) High temperature (<60°C)
Compounds Mechanism Compounds Mechanism
Additives Additives
Fluoroethylene carbonate (FEC) (ref. 59); Formation of a more conductive SEI FEC (ref. 102) Formation of a resilient SEI on
butyl sultone (BS) (ref. 60) and decrease in charge-transfer carbon anode
resistance (Rct)
Propane sulfone (PS) (ref. 103); Formation of ionically
propene sulfone (PST) (ref. 103); conducting and thermally
Tris(trimethylsilyl)borate (TMSB) stable layer on cathode
(ref. 104); 1,1’-sulfonyldiimidazole
(SDM) (ref. 105)
1,2-bis(difluoromethylsilyl) Surface protection on
ethane (FSE) (ref. 62); di-(2,2,2 both electrodes
trifluoroethyl)carbonate (DFDEC)
(ref. 106); vinylene carbonate (VC)
(ref. 107)

Lithium salts Lithium salts


Lithium tetrafluoroborate (LiBF4) (ref. 56); High ionic conductivity at low LiBOB (ref. 65); LiDFOB (ref. 98); Formation of a resilient SEI on
lithium difluoro(oxalate)borate (LiDFOB) temperatures, and decrease in Rct Li2B12F9H3 (ref. 63) carbon anode
(ref. 98)
Lithium pentafluoroethyltrifluoroborate High ionic conductivity below –10 °C Lithium tetrafluoro oxalate Surface protection on both
(LiFAB) (ref. 99) phosphate (LiTFOP) (ref. 61); lithium electrodes
(fluorosulfonyl)(nonafluorobutane)
sulfonimide (LiFNFSI) (ref. 108)

Solvents and co-solvents Solvents and co-solvents


Diethyl carbonate (DEC) (ref. 64); Decrease in melting point of Fluorinated carbonates: FEC(ref. 68); Increase in thermal stability
Ethyl methyl carbonate (EMC) (ref. 64); electrolyte F-AEC (ref. 69) towards charged electrodes
Propylene carbonate (PC) (ref. 64)
Ethyl-2,2,2-trifluoroethyl carbonate (ETFEC) Formation of a more conductive Ionic liquids (refs 46,70) Decrease in electrolyte
(ref. 100); Propyl-2,2,2-trifluoroethyl SEI and decrease in Rct Phosphates: DMMEMP (ref. 109); flammability
carbonate (PTFEC) (ref. 100) DEMEMP (ref. 109)
Ethyl butyrate (EB) (ref. 66); Formation of a more conductive
γ-butyrolactone (γBL) (ref. 101); SEI; decrease in melting point of
Methyl acetate (MA) (ref. 66); electrolyte and Rct
Ethyl acetate (EA)(ref. 66);
Ethyl propionate (EP)(ref. 66)

Effects of low temperature on cathodes. Despite the limited knowl- diffusion within the active material. The low activation energy for
edge of the exact mechanisms responsible for the decreased perfor- diffusion exhibited by certain materials can be beneficial for deploy-
mance of LIBs at low temperatures, intrinsic kinetic constraints might ment in sub-zero conditions. LVP, for instance, has an activation
partially account for the technical challenges. The charge–transfer energy of 6.57 kJmol–1, more than seven times lower than LFP,
resistance (Rct) for LFP at –20 °C has been reported to be >300% resulting in almost 300% higher relative capacity retention at –20 °C
higher than at room temperature22, invariably affecting rate perfor- (ref. 27). More generally, reducing particle size to the nanoscale,
mance. Additionally, Zhang et al. observed that the Rct for cathode and conformally coating highly conductive carbon/polymer on the
half-cells at –30 °C is highly dependent on state-of-charge, being particles’ surface can be effective in decreasing the diffusional path
much higher in the fully-lithiated state23. This resonates with the length and the inter-grain resistance, respectively. As an example,
observation by other groups that, even at moderate rates, charging Li0.99La0.01Fe0.9Mg0.1PO4/carbon aerogel composites exhibited out-
batteries at low temperatures is more challenging than discharging24. standing performance at –20 °C, displaying successful operation
Doping has been identified as an alternative to improve kinetics from 1C to 50C rate with specific capacities of 120.3 mAhg –1 and
at low temperature. Cobalt doping in lithium-rich layered materi- 56.6 mAhg–1, respectively28. Although the fabrication of such porous
als (Li1.2Ni0.2Mn0.6O2), for instance, contributes to decrease the structures is interesting to probe the intrinsic performance limits of
acti- vation energy for Ni4+ and Co4+ reduction during cell cathode materials, it drastically reduces the volumetric energy den-
discharge, improving capacity retention at –20 °C (ref. 25). sity of the cell and thus would have limited practical application.
Another example is NCM523 (Fig. 1j), in which surface
substitution with Al, even at ppm level, improved the low Anode materials
temperature performance. Selective substitution of ions, either at Negative electrodes for LIBs may present lithiation potentials below
the Li or transition metal site, have also been successful in that of reductive decomposition of organic electrolytes, giving rise
improving both ionic and electronic conductivity at extremely cold to a rich and complex interfacial chemistry. Common anode materi-
conditions26. als, such as graphite and silicon, rely on the formation of effective
Difficulties in maintaining performance below 0 °C are also asso- passivation layers to operate, and damage to these heterogeneous
ciated with high grain-boundary resistance and slow Li+ solid-state
REVIEW ARTICLE
NATURE ENERGY 2, 17108 (2017) | DOI: 10.1038/nenergy.2017.108 | www.nature.com/natureenergy
NATURE ENERGY
5

©
2
0
1
7
M
a
c
im
lla
n
P
u
b
ils
h
re
s
L
im
it
e
d
,
p
a
rt
o
fp
S
ri
n
g
re
N
a
t
u
r
e
.
A
ll
ri
g
h
ts
r
e
s
e
r
v
e
d
.
NATURE ENERGY REVIEW ARTICLE
structures renders the interfaces vulnerable at elevated tempera- improved cycle life at 55 °C (ref. 41). LTO’s performance at elevated
tures. Temperature rise also accelerates side reactions, reducing cou- temperatures is expected to be solvent-dependent, and reactiv-
lombic efficiency and cycle life1. Conversely, these layers become less ity with polymer electrolytes and ionic liquids has yet to be fully
permeable to Li+ at low temperatures, slowing down cell kinetics. explored. Reliable cycling properties demonstrated at 150 °C using
While a significant fraction of the literature on temperature effects quasi-solid electrolytes17,42 suggest that, in these cases, thermal
on cathodes focuses on the investigation of intrinsic structural sta- stability of LTO may surpass the one achieved with conventional
bility of the charged state, interfacial phenomena seems to dominate carbonate-based electrolyte.
the discussion on the anode side. At lower temperatures, the high lithiation voltage of LTO
(~1.55 V versus Li+/Li) makes it safer than graphite, due to the
Graphite. Many of the problems faced by graphite in extreme unlikeliness of lithium plating. Well-established procedures to pro-
environments can be attributed to the solid electrolyte interphase duce LTO in nanoparticle form have succeeded in retaining power
(SEI; Fig. 2a–f). Its morphology, composition and passivation prop- density below 0 °C, by decreasing the Li+ diffusion path. The lower
erties are dependent on cell temperature during both formation and activation energy for charge transfer in LTO43 is also a potential
cycling. These observations are in consonance with many reports advantage at sub-zero conditions: while graphite’s SEI has been
on SEI evolution after exposure to high temperatures, exhibiting shown to grow increasingly blocking at low temperatures33, the sur-
increased thickness and, consequently, larger cell impedance, which face layers that may exist on LTO do not seem to exhibit a distinc-
can rise by 50% upon aging at 70 °C (ref. 29). Exposition of the cell tive electrochemical signature, thus not contributing to the energy
to high temperatures for extended times leads to a decrease in the barrier for Li+ transport.
organic content in the SEI, making it more brittle30. LTO has been the material of choice for commercial batteries
The SEI might also undergo transformation due to the action of devised for applications with stringent power requirements. Granted
decomposition products from the positive electrode. For example, the low energy density is not a decisive factor, the well-known resil-
Mn-rich cathodes can suffer from metal dissolution at high tem- ience of LTO can also direct its use at temperature extremes: the
peratures, and these metal ions migrate towards the anode and pen- level of control on both particle and coating dimensions, along with
etrate the SEI, creating electronically conductive zones31. It has been the apparent inexistence of a blocking SEI, improves its potential for
proposed that Mn2+ can alter SEI composition by engaging in redox applications in demanding environments.
reactions with some of its native components, forming MnF2 and
MnCO3(ref. 32). In both scenarios, the final effect is detrimental to Silicon. Although the onset for phase transition of Li–Si alloys is
the passivation properties of the layer, leading to additional irrevers- low compared to other materials (~175–190 °C)44, it is still compat-
ible reactions with the electrolyte. ible with the range expected for most applications. When lithiated
While stabilization at elevated temperatures relies essentially Si is exposed to elevated temperatures, faster diffusion of electro-
on SEI engineering, the issues arising in colder conditions seem to lyte trapped within the SEI facilitates further reaction of Li in the
be related to intrinsic material properties. At these conditions, cell electrode, increasing the thickness of SEI and resulting in capacity
impedance rises quickly, with growing contribution from charge loss45. In contrast to graphite, the interphase layer on Si has higher
transfer and SEI to the total resistance29,33. Additionally, solid-state inorganic character, and combination of high temperatures and
diffusion of Li+ in graphite plummets below 0 °C (ref. 34). Although enormous volume changes during lithiation may lead to cracking of
there has been discussion on which of these factors becomes rate- this brittle layer. Simultaneously, the native oxide on Si surface can
limiting35, the net result is a slow-down of electrode kinetics and a react exothermically with fresh electrolyte, producing a secondary
rapid decrease in rate capability. Zhang et al. have also shown that SEI. Proper electrolyte additives have been found to produce a more
the gap between diffusion coefficients of Li+ in the charged and dis- resilient layer and improve the cell performance. Blends of ionic liq-
charged states widens at low temperatures34 (Fig. 2g). Consequently, uids with organic additives have also been shown to stabilize the SEI
ion motion is slower during lithiation, which may be partially due at temperatures as high as 100 °C (ref. 46).
to the inherent phase instability of lithiated graphite at low tem- Coating carbon on Si is another approach for enhanced capac-
peratures. The combination of all these factors also contributes to ity retention, since SEI formation is regulated by the outer car-
increased cell polarization during lithiation, favouring the deposi- bon which prevents direct contact of silicon with the electrolyte 47.
tion of metallic lithium at the anode29. Ceramic-based and metal-based coatings were also found to stabi-
Overall, graphite has exhibited fragility at both ends of the lize silicon nanostructures, as the surface layer acts as a barrier that
temperature spectrum, and applications aiming specifically for minimizes side reactions and SEI growth, and assists in maintaining
high or low temperatures could benefit from exploring other car- the electrode’s mechanical integrity upon repeated lithiation/del-
bon structures. Hard carbons, for instance, have shown promis- ithiation6. The presence of a 1.5 nm TiO 2 layer within and around
ing stability at extreme temperatures 29,36 (Fig. 2h) due to their Si nanotubes, for example, has been shown to largely improve cou-
higher resistance to solvent co-intercalation. Although the amount lombic efficiency of the cell, and to extend capacity retention by 50%
of research available for these systems is limited, alternative car- in comparison with neat silicon 48. This combination of minimized
bon materials hold potential to improve battery performance in SEI formation, decreased surface reactivity and mechanical buffer-
demanding environments. ing granted by ceramic coatings would make it a viable option for
stabilizing silicon anodes at elevated temperatures.
Spinel lithium titanium oxide. Electrochemical performance Although a recent report suggests that the impact of low operat-
of lithium titanium oxide (Li4Ti5O12, LTO) at high temperature is ing temperature on lithiation kinetics and diffusion in Si is milder
dictated by its reactivity with the electrolyte, even with its crystal than in graphite49, it is possible that the lower active material uti-
structure being stable up to 1000 °C. Despite the lack of consensus lization may increase anisotropic strain concentration, leading to
on whether and how a permanent SEI-like layer forms on LTO37, particle cracking 50. Use of silicon particles with high surface area
there is recent evidence of deposition of thick, inorganic-rich films and conductive coatings minimize some of these issues by reducing
at high temperatures, triggered by thermally activated side reac- mean Li+ diffusion path and enhancing electron transfer. However,
tions37,38. These chemical processes also result in gas evolution, experience with graphite shows that a high surface area fosters
which could damage the battery packaging 39. Coating LTO par- excessive SEI formation, which can exhibit low ionic permittivity
ticles with carbon 39, ceramics40 and polymers41 was shown to be in cold environments. Also, like graphite, deep discharge in silicon
successful in suppressing these decomposition events, leading to might show susceptibility to lithium plating.

6 NATURE ENERGY 2, 17108 (2017) | DOI: 10.1038/nenergy.2017.108 | www.nature.com/natureenergy


©
2
0
1
7
M
a
c
im
lla
n
P
u
b
ils
h
re
s
L
im
it
e
d
,
p
a
rt
o
fp
S
ri
n
g
re
N
a
t
u
r
e
.
A
ll
ri
g
h
ts
r
e
s
e
r
v
e
d
.
REVIEW ARTICLE NATURE ENERGY

a c d Electrolyte
additives
LiPF6, Rb
LiPF6, Rct
LiBF4, Rb 55 ºC
4 LiBF4, Rct

LMO cathode

Graphite anode
Log(R/K)

Mn4+
Inert SEI
Electrode structural Passivation
0 degradation
–60 –40 –20 0 20
T (oC)
b e
250

4 200
12

Capacity (mAh)
Voltage (V)

150
3

1 LiPF6, 20 ºC 100
3 4 LiPF6
2 2 LiBF4, 20 ºC
3 LiPF6, –30 ºC Li2B12F9H3/LiDFOB
4 LiBF4, –30 ºC 50 LiPF6/LiDFOB
1
0 0.3 0.6 0.9 0
Relative capacity (%) 0 300 600 900 1,200 1,500
Cycle number

Figure 3 | Tailoring salts and electrolyte additives to extend the operating temperature of Li-ion batteries. a,b, The decisive effect of charge transfer.
Although solutions of LiPF6 in alkyl-carbonates exhibit higher ionic conductivities (Rb) than when using LiBF4 (a), electrolytes containing the tetrafluoroborate
anion present lower charge-transfer resistances (Rct) throughout the entire temperature range, translating into larger capacity retention (b) at low
temperatures. c,d, Pictorial representation of dual protection activity by DFDEC and vinylene carbonate (VC) additives in Li1.13Mn0.463Ni0.203Co0.203O2/
graphite full-cells at elevated temperatures: manganese ions originating from the cathode can introduce defects at the SEI (c). The combination of these two
electrolyte additives can produce protective layers on both electrodes, suppressing metal dissolution on the cathode and forming a stable SEI on the anode
(d)106. e, Discharge capacity of Li1.1[Ni1/3Mn1/3Co1/3]0.9O2/graphite cells cycled at C/3 at 55 °C. The cyclic stability is largely enhanced using Li2B12F9H3 as a salt
and LiDFOB as additive. Panels reproduced with permission from: a,b, ref. 56 (Elsevier); e, ref. 63 (Macmillan Publishers Ltd).

The commercial success of Si anodes in bulk LIBs depend on Other materials. Several metallic-based53, intermetallic-based54
many factors, including the cost-effective fabrication of surface- and chalcogenide-based55 systems have been investigated as alter-
protected nanostructures. While the current industrial trend of native high-capacity anodes for LIBs. They all share the feature
incorporating silicon–graphite composite electrodes would ease of undergoing significant changes in particle morphology upon
some economical hurdles, it does not mitigate all challenges already lithiation, contributing to a large voltage hysteresis during cycling.
faced by graphite at extreme temperatures. Reaction kinetics are notably slow for such systems, and operation
at higher temperatures can be an effective way of reducing the volt-
Lithium metal. Metallic lithium was long forsaken in commercial age gap between charge and discharge. Conversely, although reports
devices due to low coulombic efficiency and safety concerns origi- on temperature-related effects in these materials are scarce, one can
nating from dendrite growth. Nevertheless, it has been intensively expect cell polarization to be higher at low temperatures, increasing
revisited in the past few years, for its high charge density and the the already-problematic energy inefficiency during cycling. Among
advent of other Li-based battery chemistries. the explored systems, Sn-based intermetallics, such as Cu6Sn5 and
Lithium has been largely tested at 60 °C–90 °C in combination SnSb, were observed to have greater tolerance for both thermally-
with polymer electrolytes51, showing that operation at elevated induced and electrochemically-induced lattice volume changes than
temperatures is not prohibited. More recently, it has been used at pure metals, reducing potential hysteresis53.
temperatures as high as 150 °C in LIBs containing ionic liquids 17,
highlighting its stability. Elevated temperatures have been shown Electrolyte
to improve plating/stripping efficiency and to reduce the inci- The rich variety of physical- and electro-chemical events in the elec-
dence of dendritic deposition52. While the melting point of lithium trolyte at both temperature extremes raises multifaceted challenges.
(~180 °C) imposes an intrinsic upper temperature limit for cells, At low temperatures, rapid viscosity increase has negative effects on
lithium-metal batteries would have more practical challenges in ion mobility and electrode wettability, reducing rate performance56.
the low temperature regime. Despite the limited commercial suc- Moreover, reduced conductivity causes Li+ depletion at the electrode
cess through the years, lithium metal anodes are currently used in vicinity, which may ultimately favour lithium plating29. At high tem-
electric vehicles developed by Blue Solutions both in Europe and in peratures, the electrolyte is susceptible to chemical transformations,
the United States, and the intense research activity on this topic can including disproportionation of PF6– in the presence of traces of
be projected to materialize into new ventures in the years to come. protic impurities57, and reactions with the charged electrodes58.

NATURE ENERGY 2, 17108 (2017) | DOI: 10.1038/nenergy.2017.108 | www.nature.com/natureenergy 7


©
2
0
1
7
M
a
c
im
lla
n
P
u
b
ils
h
re
s
L
im
it
e
d
,
p
a
rt
o
fp
S
ri
n
g
re
N
a
t
u
r
e
.
A
ll
ri
g
h
ts
r
e
s
e
r
v
e
d
.
NATURE ENERGY REVIEW ARTICLE
a b 100
(i)
10

0.1
100

dT/dt(°C min–1)
10 (ii)
1

0.1
100
(iii)
10

0.1
100 150 200 250 300 350
Temperature (°C)

c d e
1.2 1.2 200
Discharge capacity retention (Cx/C1)

Discharge capacity retention (Cx/C1)

150 °C
1.0 1.0
150

Capacity (mAh g–1)


0.8 0.8

0.6 0.6 100


Delithiation
0.4 0.4 Lithiation
50
25 °C 0 °C 25 °C 60 °C
0.2 0.2 C/2 rate

0.0 0.0 0
0 5 10 15 20 25 0 5 10 15 20 25 0 10 20 30 40
Cycle number Cycle number Cycle number

Figure 4 | Ionic liquids in perspective. a, Imidazolium-based ionic liquid before heating (left), and heated at 250 °C for 12 hours under vacuum (centre)
and N2 atmosphere (right). The effective thermal stability of these solvents can be much lower than suggested by thermogravimetric experiments.
b, Self-heating rate as a function of temperature for lithiated LTO in contact with EMI-FSI (i), EMI-TFSI (ii) and BMMI-TFSI (iii). The dashed lines were
obtained using ethylene carbonate, for comparison. Although certain RTIL compositions indeed increase battery safety, certain combinations can be more
reactive towards charged electrodes than carbonate-based solvents. c–d, High performance RTIL-based electrolytes: using LiTFSI/EMI-FSI (circles) as
electrolyte in graphite half-cells can provide capacities superior to LiPF6/EC+DMC (triangles) both at low (c) and high (d) temperatures. e, Cyclic stability
at 150 °C for a LTO half-cell using a hexagonal boron nitride-based quasi-solid electrolyte. These composites can support stable cell operation over a broad
temperature range, from 24 °C to 150 °C. Panels reproduced with permission from: a, ref. 80 (Elsevier); b, ref. 58 (Elsevier); c, ref. 82 (Elsevier);
d, ref. 41 (ACS); e, ref. 17 (Wiley).

Investigation of multifunctional additives and co-solvents has At high temperatures, additives such as LiTFOP (ref. 61) and
been very promising, extending LIB’s operability to both higher and FSE62 contribute to the construction of a thermally resilient SEI,
lower temperatures. There have also been intense efforts in develop- while also casting a protective coating on the cathode. The outcome
ing alternative electrolytes with enhanced thermal stability, includ- of this dual-protection mechanism (Fig. 3c,d) is the minimization of
ing ceramics and ionic liquids. The electrolyte is the sole element of structural decay on both electrodes, enhancing the cycle life when
the battery in physical contact with all other components of the cell, the battery is stored or tested at 60 °C. Many compounds initially
and the complexity of the processes it might engage in highlights the proposed as new salts, such as LiBOB and LiDFOB, have also been
challenges imposed by modern battery science. tested as additives, since their film-forming properties can benefit
the cell even at low concentrations63 (Fig. 3e). Overall, the screening
Additives and lithium salts. The efficacy of additives to enhance of new salts and additives has been very successful, and it is reason-
the battery’s thermal tolerance mainly rests on their ability to mod- able to expect it to continue having immense impact in managing
ify the surface of cathodes and anodes, and to facilitate the phase temperature-related problems in LIBs.
transfer of Li+. Recent highlights in this area are summarized in
Table 2. Since ion transport through SEI slows down below room Organic solvents and co-solvents. The battery community has long
temperature, the formation of more compact and conductive pas- realized the usefulness of combining different co-solvents to add
sivation films, as provided by FEC 59 and butyl sultone60, is highly functionalities to electrolytes (Table 2). A common approach for low
beneficial to rate performance. Optimizing Rcthas also been shown temperature applications is the addition of low-melting-point, low-
to be effective: although electrolytes containing LiBF4 exhibit lower viscosity solvents, such as EMC and PC64, to extend the electrolyte’s
ionic conductivities than with LiPF6 (Fig. 3a), the lower Rct achieved liquid range and assure good ionic conductivity. Recently, this func-
with tetrafluoroborate anions lead to better capacity retention tion has been further combined with the ability to produce a more
at –30 °C (Fig. 3b)56. conductive SEI and to decrease Rct, for which γ-butyrolactone65 and

8 NATURE ENERGY 2, 17108 (2017) | DOI: 10.1038/nenergy.2017.108 | www.nature.com/natureenergy


©
2
0
1
7
M
a
c
im
lla
n
P
u
b
ils
h
re
s
L
im
it
e
d
,
p
a
rt
o
fp
S
ri
n
g
re
N
a
t
u
r
e
.
A
ll
ri
g
h
ts
r
e
s
e
r
v
e
d
.
REVIEW ARTICLE NATURE ENERGY

a b
LiTFSI LiPF 6 LiBF4
14
0.5 LiTFSI
55 °C 85 °C 85 °C LiBETI
12
Specific current (mA cm–2)

Current density (mA cm–2)


0.0 LiPF6
10
LiBF4
1.5 100 °C 1st 8
2nd
1.0 3rd 6

4
0.5 65 °C
100 °C 2
0.0
0
3 4 5 3 4 5 3 4 5 40 60 80 100 120
Potential (V) versus Li/Li + Temperature (°C)

Figure 5 | Aluminium corrosion in a pyrrolidinium-based ionic liquid at elevated temperatures. a, Cyclic voltammograms of Al/Li cells containing
different electrolytes. The passivation properties are highly dependent on temperature. b, Linear sweep thermometry obtained with the use of different
lithium salts, and voltage held at 5.0 V. The onset for aluminium corrosion occurred at different temperatures, being minimum for LiTFSI and maximum for
LiBF4. Figure 5 reproduced with permission from ref. 93 (Electrochemical Society).

short-chain acetates66 have been quite successful. Smart et al.67 have The thermal and chemical stability of ceramics has also spurred
also shown that addition of 20vol% methyl propionate or ethyl pro- interest in utilizing crystalline ion conductors as electrolytes. Oxides
pionate to EC:DMC mixtures provides over 70% capacity retention and phosphates are natural candidates for enabling safe operation
when the cell operates at –60 °C. of LIBs in extreme environments, due to their intrinsic structural
On the high-temperature end, the strategy has been focused stability at high temperatures75. Recently, LiPON has been shown
on reducing electrolyte flammability and improving its stability to be stable up to 200 °C (ref. 76), although the cell exhibited less
towards charged electrodes. Much work has been done in explor- capacity than at lower temperatures and long-term performance
ing fluorinated carbonates as co-solvents, since the addition of fluo- was not reported. The sulfide glass 95(0.7Li2S–0.3P2S5)–5Li3PO4 is
rine moieties increases thermal resilience and largely improves the also known to provide stable cycling at 100 °C (ref. 77), despite the
anodic stability of these compounds. Consequently, solvents such questionable environmental inertness of this family of conductors78.
as FEC68 and F-AEC69 have been found to drastically improve safety Although the resilience of ceramics is not to be disputed, the ability
and cycle life at 55 °C. Ionic liquids have also been investigated of these rather brittle materials to sustain thermal cycling in a bat-
as co-solvents, usually in combination with PC, to reduce flam- tery has yet to be confirmed.
mability70 and provide additional building blocks for a thermally Reports on the performance of solid electrolytes at low tem-
resilient SEI46. peratures are scarce in the literature. It can be anticipated that the
Based on the current state of the literature, it is realistic to pro- large interfacial resistances79 existing in most all-solid-state cells are
ject that these solvents could enable satisfactory battery operation a major challenge for retaining power capability in cold environ-
under very cold environments. However, intrinsic limitations on ments. The ionic conductivities, already low at room temperature,
their physical properties — such as flammability, high vapour pres- also plunge as temperature decreases. The transition to solid elec-
sure and low flash point — limit the ability of these compositions to trolytes is a possible way for LIBs to move forward, but the variety of
sustain long term storage and cycling beyond 80 °C. challenges to be solved to achieve peak performance, even at room
temperature, makes the complete substitution of liquid electrolytes
Solid electrolytes. The absence of volatile or flammable compounds a task hard to undertake soon.
is expected to make solid electrolytes safer than their liquid coun-
terparts at elevated temperatures. Besides thermal stability, solid Room-temperature ionic liquids. These hold the promise of pro-
polymer electrolytes (SPEs) and ceramic conductors offer attractive ducing safer devices and enabling battery operation at very high
design possibilities, enabling thin film batteries and easy integration temperatures. The main appeal of these compounds is their non-
into microdevices. flammability and negligible vapour pressure within a wide tempera-
Although electrochemical testing beyond 60 °C is expedient to ture range17. Early assessment of their thermal stability neglected
probe properties of SPEs, these materials might present issues at ageing effects on the structural integrity. Albeit not producing
temperature extremes. For example, slow recrystallization kinetics of significant amounts of volatile products until ~300 °C, it has been
the ubiquitous poly(ethylene oxide) (PEO) renders ionic conductiv- shown that partial decomposition of the cation can occur after
ity highly dependent on the thermal history of the films71. Recently, extended heat treatments at temperatures as low as 140 °C (ref. 80)
adoption of porous structures of thermally stable polyimide 72 and (Fig. 4a). Another important point is that their effective thermal sta-
poly(tetrafluoroethylene) 73 as skeletons for PEO-succinonitrile- bility when in contact with charged electrodes might divert from the
LiBOB composites provided good results, with conductivities stable inertness displayed by the pure solvents. Wang et al. (ref. 58) have
for at least 120 h at 160 °C. Unfortunately, these latter studies did not shown that, for certain room-temperature ionic liquids (RTILs),
present further testing on a cell level, and the actual performance onset for thermal runaway can actually be lower than for carbonate-
is unknown. Notwithstanding the uncertainties in SPE behaviour based electrolytes (Fig. 4b). Although this property varies widely
beyond 100 °C, these materials have been successful in extending with RTIL structure, TFSI anions and some onium cations are in
battery operation to 90 °C (refs 51,74). Recent commercialization fact more stable than organic carbonates.
of such SPE-based batteries was implemented in the Blue Car/Bus Besides the elevated cost of high-purity RTILs, they are down-
project by the Bolloré Group, wherein battery packs are heated and played for their performance at room temperatures and lower, due
operated between 60–80 °C. to their high viscosity, large R ct (ref. 46) and low Li+ transference

NATURE ENERGY 2, 17108 (2017) | DOI: 10.1038/nenergy.2017.108 | www.nature.com/natureenergy 9


©
2
0
1
7
M
a
c
im
lla
n
P
u
b
ils
h
re
s
L
im
it
e
d
,
p
a
rt
o
fp
S
ri
n
g
re
N
a
t
u
r
e
.
A
ll
ri
g
h
ts
r
e
s
e
r
v
e
d
.
NATURE ENERGY REVIEW ARTICLE
a
750

700

650

Na-S

Energy density (Wh kg-1)


150
Li-ion

Li-ion polymer

Molten
100
salt
Ni-MH

Li-metal
polymer
50
Lead acid

0
–40 –20 0 20 40 60 80 100 300 400
Temperature (ºC)
b
Consumer electronics Military Liquid Solid ceramic Solid polymer
300

PEO-polyimide (ref. 72)


LiPON (ref. 76)
es (ref. 116)
250
With molten salt (ref. 20)

With RTIL (ref. 17)

PEO-PTFE (ref. 73)


With RTIL (ref. 46)
Melting point
3D carbon decorated (ref. 114)

PP13-TFSI (ref. 17)


Polymer electrolyte (ref. 51)
Organic electrolyte (ref. 97)

luorinated carbonates (ref. 69)


Mixture of carbonates (ref. 64)
With RTIL (ref. 17)

With LiBOB (ref. 94)


With LiPF6 salt (ref. 112)

Carbon-free electrode (ref. 115)

Amorphous monolithic (ref. 49)


Full cell with LMO (ref. 113)

200 With PC-RTIL (ref. 46)


Commercial grade (ref. 111)

77)
Carbon coated (ref. 27)
AlPO4 coated (ref. 7)

With RTIL (ref. 82)

Graphite mixture (ref. 29)


Hard carbon (ref. 36)

Lactones (ref. 101)


oated (ref. 97)
ed (ref. 94)

Ph
With FEC (ref. 68)

30)
With LTO (ref. 113)

With VC (ref. 23)


With LTO (ref. 9)

150 os .

EMI-FSI (ref. 82)


ef

PEO (ref. 51)


Esters (ref. 67)

(r
Temperature (ºC)

ph
oa
tc c PV
at
100 O DF
T
L .
ef
(r
Ca F s
rb
glas
50 on
Sulfide

–50

Cathodes Anodes Electrolytes

Figure 6 | Reported thermal limits for Li-ion battery materials and rechargeable battery systems. a, Typical energy densities and temperature windows of
commercially available rechargeable batteries. b, Summary of the operating temperatures achieved with different cathode, anode and electrolyte materials.
The diversity of LIB-compatible compositions is an asset to the development of thermally resilient devices. Data are from refs 7, 9, 10, 12, 14, 15, 17, 28, 31,
35, 36, 41, 51, 54, 56, 70, 73–75, 80, 89, 90, 93, 94, 100, 111–116 as indicated in the figure.

number17,46,81. Strategies to enable performance at lower tempera- However, practical battery operation in such cold environments has
tures involve optimizing charge-transfer kinetics46,81. Recent devel- yet to be demonstrated.
opment of highly conductive FSI-based RTILs has also improved At the high-temperature end, exploitation of the possibilities ena-
rate performance (Fig. 4c,d), and addition of LiBOB was found bled by RTILs are limited by the lack of state-of-the-art thermally
to be especially effective at improving capacity retention at 0 °C stable separators. Although RTIL-based polymer gels have been
(ref. 82). Although many RTILs freeze below –10 °C, operability at explored85, rapid decay of mechanical properties with temperature
lower temperatures can be explored by forming eutectic mixtures do little to extend the operating range. Performance in more extreme
with suitable combinations of ionic liquids and lithium salts83,84. conditions has been recently demonstrated by immobilizing RTILs

10 NATURE ENERGY 2, 17108 (2017) | DOI: 10.1038/nenergy.2017.108 | www.nature.com/natureenergy


©
2
0
1
7
M
a
c
im
lla
n
P
u
b
ils
h
re
s
L
im
it
e
d
,
p
a
rt
o
fp
S
ri
n
g
re
N
a
t
u
r
e
.
A
ll
ri
g
h
ts
r
e
s
e
r
v
e
d
.
REVIEW ARTICLE NATURE ENERGY

in ceramic materials, as natural clay42 and hexagonal boron nitride17 directly transferred to alkyl-carbonate systems, it serves the pur-
(h-BN). The combination of h-BN and a piperidinium-based RTIL pose of raising awareness of the fact that passivation behaviour can
enabled LTO half-cells to operate in the 24–150 °C range, with good be very dynamic.
capacity retention and rate capability during extended cycling at
high temperatures17 (Fig. 4e). Despite the many exciting develop- Outlook
ments during the last decade, RTILs share a common challenge with Battery science has always been a game of trade-offs, and this
SPEs: compositions suitable for deployment at high temperatures also holds when considering thermal factors. Figure 6a compares
often underperform in milder environments, restricting them to energy density of commercial rechargeable batteries against their
niche applications. operable temperature window. Despite LIB’s clear prominence,
the environmental frontiers for these batteries are still modest
Separators, binders and current collectors compared to the possibilities offered by its core materials. Many
Although temperature effects are less studied in these electrochemi- reports focus on extending the thermal usability of individual
cally inactive components, their state-of-health is essential to sus- components, and Fig. 6b maps some of the standard and prom-
tain cell operation. ising materials against their temperature limits, each specifically
discussed in this review.
Separators. Serving as a physical barrier between electrodes, tradi- In designing an optimum battery, materials selection is obvi-
tional polyolefin-based separators also feature a ‘shut-down’ mecha- ously made with compatibility in mind. Apart from immediate
nism when exposed to overheating and/or overcharge. A practical reactivity, slow transformations should be considered before suc-
approach to enhance their thermal stability is to coat inorganic nan- cessful implementation, and being mindful of the interfacial reac-
oparticles in the separator films. Several studies encompassed the tions between electrolytes and electrodes at every state-of-charge
use of ceramic materials, such as SiO2 and Al2O3, to enhance surface is essential. Fluorination of alkyl-carbonates enhances the compat-
morphology, reducing thermal shrinkage at elevated temperatures ibility between electrolyte and charged electrodes, whereas salts
and improving ionic conductivities86. such as LiBOB and LiDFOB confer various benefits to the thermal
Thermosetting polymers have also been investigated as alterna- resilience of the cell. Phosphate-based cathodes are particularly
tive separators. The thermal and chemical stability of polyimides, stable at high temperatures, with LiFePO 4 already transitioning
especially when cross-linked into three-dimensional networks, into the consumer’s market. LVP (Li3V2(PO4)3) is another interest-
makes them especially attractive for high-temperature applications. ing material to be watched, since it inherits the sturdiness from
Jiang et al. demonstrated that electrospun nano-fibrous polyim- phosphates while also operating at higher voltages than its ferrous
ide membranes exhibited no thermal shrinkage at 150 °C (ref. 87). relative and presenting better low-temperature performance.
Lee etal. havealsoreported thatcoating polyimide fiberswithalumina On the other hand, reactions at the anode are highly dependent
is beneficial for the separator’s wetting and transport properties88. on the quality and quantity of the SEI formed. The usability of a bat-
tery is dictated by the nature and evolution of this passivation layer
Binders. Mechanical properties of commonly used poly(vinylene under the operating temperature scenarios. Li + transport through
fluoride) (PVDF) are compromised at elevated temperatures, as SEI is one of the major limiting factors at low temperatures, and
the polymer migrates to the electrode surface at 120 °C, affect- eventually favours lithium plating during cell charging. The choice
ing structural integrity 30,37. Yan et al. demonstrated that polyim- of cathode could also influence the anode passivation, due to the
ide can be synthesized via a one-pot solution polycondensation, presence of minute dissolved ions from its structure. Formation
and incorporated into a graphite anode 89. Such cell exhibited rate protocols, which include pretreatments and use of specific addi-
capability and cycling behaviour is superior to that of conventional tives, have been a closely guarded know-how of the industry,
binders. Despite the electrochemical characterization of this work and it is uncertain how they will adapt to accommodate thermal
being carried out at room temperature, the use of this thermally extremes. Diagnostic analyses are essential to rationalize perfor-
stable polymer lends itself as an attractive binder candidate for mance decay, while systematic studies on the correlation between
extreme environments. thermal history and state-of-health are necessary to project long-
term behaviour. While traditional graphite electrodes have a high
Current collector. Copper has been suggested to be susceptible to degree of uncertainty at either of the thermal extremes, LTO, hard
oxidation by impurities in the electrolyte90, and this process can be carbons and metallic lithium seem to be promising alternatives.
expected to be accelerated beyond 50 °C, both by enhanced kinet- The collective knowledge currently available on the subtleties of
ics and decay of LiPF6 into HF57. Although the anode current col- LIBs is astounding, with an ever-growing library of materials being
lector (CC) has not received much attention, aluminium has been incorporated to the field. However, commercial adoption of these
deeply investigated, with the main goal of supporting high voltage materials to engineer battery packs with thermal tolerance has a
cathodes. Little attention, however, has been paid to temperature- long way to go. While most of these reports are impressive and pre-
related effects on the passivation properties of this interface. sent valuable insights, they are usually focused on a single compo-
A possible problem at elevated temperatures is the loss of pas- nent of a battery. In addition to the material properties, cell-level
sivation properties due to cracking and dissolution of its com- performance is governed by secondary conditions and interactions
ponents. Intrinsic electrolyte limitations have made it difficult to between its components. Industrial know-how clearly surpasses
probe specific effects on the CC, especially since some thermally open-sourced scientific reports in this area, and it is yet to be seen
stable salts, such as LiTFSI, seem to be unable to protect alumin- how technology transitions between them.
ium beyond 3.7 V versus Li+/Li (ref. 91). Nevertheless, carbonate-
based electrolytes containing LiTFSI-LiBOB mixtures exhibited Received 2 November 2016; accepted 4 June 2017;
promising passivating behaviour up to 60 °C (ref. 92). published 24 July 2017
The lack of volatility of RTILs make them ideal systems to probe
temperature-related phenomena on CCs. As shown in Fig. 5, References
anodic stability of aluminium in a pyrrolidinium-based RTIL 1. Tarascon, J. M. & Armand, M. Issues and challenges facing rechargeable
lithium batteries. Nature 414, 359–367 (2001).
varies widely both with the salt of choice and the operating tem- 2. Forgez, C., Vinh Do, D., Friedrich, G., Morcrette, M. & Delacourt, C.
perature93. A highlight is LiBF4, which seems able to support 5 V Thermal modeling of a cylindrical LiFePO4/graphite lithium-ion battery.
chemistries beyond 85 °C. Although this knowledge cannot be J. Power Sour. 195, 2961–2968 (2010).

NATURE ENERGY 2, 17108 (2017) | DOI: 10.1038/nenergy.2017.108 | www.nature.com/natureenergy 11


©
2
0
1
7
M
a
c
im
lla
n
P
u
b
ils
h
re
s
L
im
it
e
d
,
p
a
rt
o
fp
S
ri
n
g
re
N
a
t
u
r
e
.
A
ll
ri
g
h
ts
r
e
s
e
r
v
e
d
.
NATURE ENERGY REVIEW ARTICLE
3. Feng, X. et al. Thermal runaway features of large format prismatic lithium ion 28. Zhang, H., Xu, Y., Zhao, C., Yang, X. & Jiang, Q. Effects of carbon coating and
battery using extended volume accelerating rate calorimetry. J. Power Sour. metal ions doping on low temperature electrochemical properties of LiFePO4
255, 294–301 (2014). cathode material. Electrochimica Acta 83, 341–347 (2012).
4. Williard, N., He, W., Hendricks, C. & Pecht, M. Lessons learned from the 787 29. Waldmann, T., Wilka, M., Kasper, M., Fleischhammer, M. &
Dreamliner issue on lithium-ion battery reliability. Energies Wohlfahrt-Mehrens, M. Temperature dependent ageing mechanisms in lithium-
6, 4682–4695 (2013). ion batteries — a post-mortem study. J. Power Sour.
5. Linden, D. R. & Thomas B. Handbook of Batteries. 3rd edn 262, 129–135 (2014).
(McGraw-Hill, 2002). 30. Bodenes, L. et al. Lithium secondary batteries working at very high
6. Mauger, A. & Julien, C. Surface modifications of electrode materials for temperature: capacity fade and understanding of aging mechanisms.
lithium-ion batteries: status and trends. Ionics 20, 751–787 (2014). J. Power Sour. 236, 265–275 (2013).
7. Cho, Y., Eom, J. & Cho, J. High performance LiCoO2 cathode materials at 31. Jalkanen, K. et al. Cycle aging of commercial NMC/graphite pouch cells at
60 °C for lithium secondary batteries prepared by the facile nanoscale different temperatures. Appl. Energy 154, 160–172 (2015).
dry-coating method. J. Electrochem. Soc. 157, A617–A624 (2010). 32. Shin, H., Park, J., Sastry, A. M. & Lu, W. Degradation of the solid electrolyte
8. Aurbach, D. et al. Review on electrode–electrolyte solution interactions, interphase induced by the deposition of manganese ions. J. Power Sour.
related to cathode materials for Li-ion batteries. J. Power Sour. 284, 416–427 (2015).
165, 491–499 (2007). 33. Zhang, S. S., Xu, K. & Jow, T. R. Electrochemical impedance study on the low
9. Kong, D. P., Ping, P., Wang, Q. S. & Sun, J. H. Study on high temperature of Li-ion batteries. Electrochimica Acta
temperature stability of LiNi0.33Co0.33Mn0.33O2/Li4Ti5O12 cells from the safety 49, 1057–1061 (2004).
perspective. J. Electrochem. Soc. 163, A1697–A1704 (2016). 34. Zhang, S. S., Xu, K. & Jow, T. R. Low temperature performance of graphite
10. Yan, C., Xu, Y., Xia, J., Gong, C. & Chen, K. Tris(trimethylsilyl) borate electrode in Li-ion cells. Electrochimica Acta 48, 241–246 (2002).
as an electrolyte additive for high-voltage lithium-ion batteries using 35. Huang, C. K., Sakamoto, J. S., Wolfenstine, J. & Surampudi, S. The limits of
LiNi1/3Mn1/3Co1/3O2 cathode. J. Energy Chem. 25, 659–666 (2016). low-temperature performance of Li-ion cells. J. Electrochem. Soc.
11. Ma, L., Nie, M., Xia, J. & Dahn, J. A systematic study on the reactivity of 147, 2893–2896 (2000).
different grades of charged Li[NixMnyCoz]O2 with electrolyte at elevated 36. Zheng, H., Qu, Q., Zhang, L., Liu, G. & Battaglia, V. S. Hard carbon: a
temperatures using accelerating rate calorimetry. J. Power Sour. promising lithium-ion battery anode for high temperature applications with
327, 145–150 (2016). ionic electrolyte. Roy. Soc. Chem. Adv. 2, 4904–4912 (2012).
12. Bloom, I. et al. Effect of cathode composition on capacity fade, impedance rise 37. Gieu, J. B., Courrèges, C., El Ouatani, L., Tessier, C. & Martinez, H.
and power fade in high-power, lithium-ion cells. J. Power Sour. Temperature effects on Li4Ti5O12 electrode/electrolyte interfaces at the first
124, 538–550 (2003). cycle: a X-ray photoelectron spectroscopy and scanning auger microscopy
13. Nitta, N., Wu, F., Lee, J. T. & Yushin, G. Li-ion battery materials: present and study. J. Power Sour. 318, 291–301 (2016).
future. Mater. Today 18, 252–264 (2015). 38. Nordh, T., Younesi, R., Brandell, D. & Edström, K. Depth profiling the solid
14. Huang, Y. et al. Thermal stability and reactivity of cathode materials for Li-ion electrolyte interphase on lithium titanate (Li4Ti5O12) using synchrotron-based
batteries. ACS Appl. Mater. Interfaces 8, 7013–7021 (2016). photoelectron spectroscopy. J. Power Sour. 294, 173–179 (2015).
The authors investigate the thermal stability of several cathode materials in 39. He, Y. B. et al. Gassing in Li4Ti5O12-based batteries and its remedy. Sci. Rep.
the delithated state, showing that layered cathodes are the least stable, and 2, 913 (2012).
phosphates the most. 40. Han, C. et al. Suppression of interfacial reactions between Li4Ti5O12 electrode
15. Zhao, H. et al. Enhanced elevated-temperature performance of and electrolyte solution via zinc oxide coating. Electrochimica Acta
LiAlxSi0.05Mg0.05Mn1.90–xO4 (0 ≤ x ≤ 0.08) cathode materials for high- 157, 266–273 (2015).
performance lithium-ion batteries. Electrochimica Acta 199, 18–26 (2016). 41. Lu, Q. et al. A polyimide ion-conductive protection layer to suppress side
16. Sun, Y. K. et al. Nanostructured high-energy cathode materials for advanced reactions on Li4Ti5O12 electrodes at elevated temperature. Roy. Soc. Chem. Adv.
lithium batteries. Nat. Mater. 11, 942–947 (2012). 4, 10280 (2014).
17. Rodrigues, M.-T. F. et al. Hexagonal boron nitride-based electrolyte composite 42. Kalaga, K. et al. Quasi-solid electrolytes for high temperature lithium ion
for Li-ion battery operation from room temperature to 150 °C. batteries. ACS Appl. Mater. Interfaces 7, 25777–25783 (2015).
Adv. Energy Mater. 6, 1600218 (2016). 43. Xu, K. & von Wald Cresce, A. Li+-solvation/desolvation dictates interphasial
This work demonstrates that ionic liquids can allow Li-ion batteries to processes on graphitic anode in Li ion cells. J. Mater. Res.
operate from room temperature to at least 150 °C, with relatively high 27, 2327–2341 (2012).
coulombic efficiencies. 44. Wang, Y. & Dahn, J. Phase changes in electrochemically lithiated silicon at
18. Hagh, N. M. & Amatucci, G. G. Effect of cation and anion doping on elevated temperature. J. Electrochem. Soc. 153, A2314 (2006).
microstructure and electrochemical properties of the LiMn 1.5Ni0.5O4−δ spinel. 45. Fan, H. et al. Electrochemical properties and thermal stability of silicon
J. Power Sour. 256, 457–469 (2014). monoxide anode for rechargeable lithium-ion batteries. Electrochemistry
19. Maccario, M., Croguennec, L., Le Cras, F. & Delmas, C. Electrochemical 84, 574–577 (2016).
performances in temperature for a C-containing LiFePO4 composite 46. Ababtain, K. et al. Ionic liquid-organic carbonate electrolyte blends to stabilize
synthesized at high temperature. J. Power Sour. 183, 411–417 (2008). silicon electrodes for extending lithium ion battery operability to 100 °C.
20. Muñoz-Rojas, D. et al. Development and implementation of a high ACS Appl. Mater. Interfaces 8, 15242–15249 (2016).
temperature electrochemical cell for lithium batteries. Electrochem. Commun. 47. Profatilova, I. A. et al. Thermally induced reactions between lithiated nano-
9, 708–712 (2007). silicon electrode and electrolyte for lithium-ion batteries. J. Electrochem. Soc.
21. Chang, H.-H. et al. Effects of TiO2 coating on high-temperature cycle 159, A657–A663 (2012).
performance of LiFePO4-based lithium-ion batteries. J. Power Sour. 48. Lotfabad, E. M. et al. Si nanotubes ALD coated with TiO2, TiN or Al2O3 as
185, 466–472 (2008). high performance lithium ion battery anodes. J. Mater. Chem. A
22. Gao, F. & Tang, Z. Kinetic behavior of LiFePO4/C cathode material for 2, 2504–2516 (2014).
lithium-ion batteries. Electrochimica Acta 53, 5071–5075 (2008). 49. Markevich, E., Salitra, G. & Aurbach, D. Low temperature performance of
23. Zhang, S. S., Xu, K. & Jow, T. R. The low temperature performance of Li-ion amorphous monolithic silicon anodes: comparative study of silicon and
batteries. J. Power Sour. 115, 137–140 (2003). graphite electrodes. J. Electrochem. Soc. 163, A2407–A2412 (2016).
24. Fan, J. & Tan, S. Studies on charging lithium-ion cells at low temperatures. 50. An, Y. et al. Mitigating mechanical failure of crystalline silicon electrodes for
J. Electrochem. Soc. 153, A1081–A1092 (2006). lithium batteries by morphological design. Phys. Chem. Chem. Phys.
25. Kou, J. et al. Role of cobalt content in improving the low-temperature 17, 17718–17728 (2015).
performance of layered lithium-rich cathode materials for lithium-ion 51. Croce, F., Sacchetti, S. & Scrosati, B. Advanced, lithium batteries based on
batteries. ACS Appl. Mater. Interfaces 7, 17910–17918 (2015). high-performance composite polymer electrolytes. J. Power Sour.
26. Li, G. et al. Effect of trace Al surface doping on the structure, surface 162, 685–689 (2006).
chemistry and low temperature performance of LiNi0.5Co0.2Mn0.3O2 cathode. 52. Akolkar, R. Modeling dendrite growth during lithium electrodeposition
Electrochimica Acta 212, 399–407 (2016). at sub-ambient temperature. J. Power Sour. 246, 84–89 (2014).
27. Rui, X. H., Jin, Y., Feng, X. Y., Zhang, L. C. & Chen, C. H. A comparative 53. Winter, M. & Besenhard, J. O. Electrochemical lithiation of tin and tin-based
study on the low-temperature performance of LiFePO4/C and Li3V2(PO4)3/C intermetallics and composites. Electrochimica Acta
cathodes for lithium-ion batteries. J. Power Sour. 196, 2109–2114 (2011). 45, 31–50 (1999).
This work provides a detailed investigation on the origins of the enhanced 54. Jansen, A. N., Clevenger, J. A., Baebler, A. M. & Vaughey, J. T. Variable
performance of LVP at –20 °C, highlighting the benefit of a low activation temperature performance of intermetallic lithium-ion battery anode materials.
energy for Li+ diffusion in the lattice to capacity retention. J. Alloys Compounds 509, 4457–4461 (2011).

12 NATURE ENERGY 2, 17108 (2017) | DOI: 10.1038/nenergy.2017.108 | www.nature.com/natureenergy


©
2
0
1
7
M
a
c
im
lla
n
P
u
b
ils
h
re
s
L
im
it
e
d
,
p
a
rt
o
fp
S
ri
n
g
re
N
a
t
u
r
e
.
A
ll
ri
g
h
ts
r
e
s
e
r
v
e
d
.
REVIEW ARTICLE NATURE ENERGY

55. Sen, U. K. & Mitra, S. High-rate and high-energy-density lithium-ion battery 81. Rodrigues, M.-T. F., Lin, X., Gullapalli, H., Grinstaff, M. W. & Ajayan, P. M.
anode containing 2D MoS2 nanowall and cellulose binder. Rate limiting activity of charge transfer during lithiation from ionic liquids.
ACS Appl. Mater. Interfaces 5, 1240–1247 (2013). J. Power Sour. 330, 84–91 (2016).
56. Zhang, S. S., Xu, K. & Jow, T. R. A new approach toward improved low 82. Yamagata, M. et al. High-performance graphite negative electrode in a
temperature performance of Li-ion battery. Electrochem. Commun. bis(fluorosulfonyl)imide-based ionic liquid. J. Power Sour.
4, 928–932 (2002). 227, 60–64 (2013).
57. Campion, C. L., Li, W. & Lucht, B. L. Thermal decomposition of LiPF6-based 83. Zhou, Q. et al. Phase behavior of ionic liquid–LiX mixtures: pyrrolidinium
electrolytes for lithium-ion batteries. J. Electrochem. Soc. cations and TFSI–anions — linking structure to transport properties.
152, A2327 (2005). Chem. Mater. 23, 4331–4337 (2011).
58. Wang, Y. et al. Accelerating rate calorimetry studies of the reactions 84. Tsai, W.-Y. et al. Outstanding performance of activated graphene based
between ionic liquids and charged lithium ion battery electrode materials. supercapacitors in ionic liquid electrolyte from −50 to 80 °C. Nano Energy
Electrochimica Acta 52, 6346–6352 (2007). 2, 403–411 (2013).
This work disputes some claims of thermal stability of ionic liquids, 85. Lee, J. H. et al. Hybrid ionogel electrolytes for high temperature lithium
showing that some compositions undergo thermal runaway at lower batteries. J. Mater. Chem. A 3, 2226–2233 (2015).
temperatures than carbonate-based solvents. 86. Jeong, H.-S., Hong, S. C. & Lee, S.-Y. Effect of microporous structure on
59. Liao, L. et al. Fluoroethylene carbonate as electrolyte additive to improve low thermal shrinkage and electrochemical performance of Al2O3/poly(vinylidene
temperature performance of LiFePO4 electrode. Electrochimica Acta fluoride-hexafluoropropylene) composite separators for lithium-ion batteries.
87, 466–472 (2013). J. Membrane Sci. 364, 177–182 (2010).
60. Liao, L. et al. Enhancement of low-temperature performance of LiFePO4 87. Jiang, W. et al. A high temperature operating nanofibrous polyimide separator
electrode by butyl sultone as electrolyte additive. Solid State Ionics in Li-ion battery. Solid State Ionics 232, 44–48 (2013).
254, 27–31 (2014). 88. Lee, J., Lee, C.-L., Park, K. & Kim, I.-D. Synthesis of an Al2O3-coated
61. Qin, Y., Chen, Z., Liu, J. & Amine, K. Lithium tetrafluoro oxalato phosphate as polyimide nanofiber mat and its electrochemical characteristics as a separator
electrolyte additive for lithium-ion cells. Electrochem. Solid-State Lett. for lithium ion batteries. J. Power Sour. 248, 1211–1217 (2014).
13, A11 (2010). 89. Yan, X. et al. Polyimide binder by combining with polyimide separator
62. Yamagiwa, K. et al. Improved high-temperature performance and surface for enhancing the electrochemical performance of lithium ion batteries.
chemistry of graphite/LiMn2O4 Li-ion cells by fluorosilane-based electrolyte Electrochimica Acta 216, 1–7 (2016).
additive. Electrochimica Acta 160, 347–356 (2015). 90. Zhao, M. et al. Electrochemical stability of copper in lithium-ion battery
63. Chen, Z. et al. New class of nonaqueous electrolytes for long-life and safe electrolytes. J. Electrochem. Soc. 147, 2874 (2000).
lithium-ion batteries. Nat. Commun. 4, 1513 (2013). 91. Morita, M., Shibata, T., Yoshimoto, N. & Ishikawa, M. Anodic behavior
Proposes Li2B12F9H3 as a new Li-salt, exhibiting enhanced stability towards of aluminum current collector in LiTFSI solutions with different solvent
charged electrodes at elevated temperatures and overcharge protection. compositions. J. Power Sour. 119–121, 784–788 (2003).
64. Yaakov, D., Gofer, Y., Aurbach, D. & Halalay, I. C. On the study of electrolyte 92. Chen, X. et al. Mixed salts of LiTFSI and LiBOB for stable LiFePO4-based
solutions for Li-ion batteries that can work over a wide temperature range. batteries at elevated temperatures. J. Mater. Chem. A 2, 2346 (2014).
J. Electrochem. Soc. 157, A1383 (2010). 93. Mun, J. et al. Linear-sweep thermammetry study on corrosion behavior of Al
65. Xu, K. Tailoring electrolyte composition for LiBOB. J. Electrochem. Soc. current collector in ionic liquid solvent. Electrochem. Solid-State Lett.
155, A733 (2008). 13, A109 (2010).
66. Smart, M. C., Ratnakumar, B. V. & Surampudi, S. Use of organic esters 94. Li, J., Zhu, Y., Wang, L. & Cao, C. Lithium titanate epitaxial coating on spinel
as cosolvents in electrolytes for lithium-ion batteries with improved low lithium manganese oxide surface for improving the performance of lithium
temperature performance. J. Electrochem. Soc. 149, A361 (2002). storage capability. ACS Appl. Mater. Interfaces 6, 18742–18750 (2014).
A comprehensive analysis of the implementation of acetates as co-solvents 95. Zhang, J. et al. Sustainable, heat-resistant and flame-retardant cellulose-based
in electrolytes, extending battery operability to –70 °C. composite separator for high-performance lithium ion battery. Sci. Rep.
67. Smart, M. C., Ratnakumar, B. V., Chin, K. B. & Whitcanack, L. D. Lithium- 4, 3935 (2014).
ion electrolytes containing ester cosolvents for improved low temperature 96. Zheng, F. et al. Surfactants assisted synthesis and electrochemical properties of
performance. J. Electrochem. Soc. 157, A1361 (2010). nano-LiFePO4/C cathode materials for low temperature applications.
68. Hu, L., Zhang, Z. & Amine, K. Fluorinated electrolytes for Li-ion battery: J. Power Sour. 288, 337–344 (2015).
an FEC-based electrolyte for high voltage LiNi0.5Mn1.5O4/graphite couple. 97. Liao, L. et al. Effects of temperature on charge/discharge behaviors of LiFePO4
Electrochem. Commun. 35, 76–79 (2013). cathode for Li-ion batteries. Electrochimica Acta
69. Zhang, Z. et al. Fluorinated electrolytes for 5 V lithium-ion battery chemistry. 60, 269–273 (2012).
Energy Environ. Sci. 6, 1806 (2013). 98. Shui Zhang, S. An unique lithium salt for the improved electrolyte of Li-ion
70. Kühnel, R.-S. & Balducci, A. Lithium ion transport and solvation in N-Butyl- battery. Electrochem. Commun. 8, 1423–1428 (2006).
N-methylpyrrolidinium Bis(trifluoromethanesulfonyl)imide–Propylene 99. Zhou, Z.-B., Takeda, M., Fujii, T. & Ue, M. Li [C2F5BF3] as an electrolyte salt
carbonate mixtures. J. Phys. Chem. C 118, 5742–5748 (2014). for 4 V class lithium-ion cells. J. Electrochem. Soc.
71. Choi, B. K. & Kim, Y. W. Thermal history effects on the ionic conductivity of 152, A351–A356 (2005).
PEO-salt electrolytes. Mater. Sci. Eng. B 107, 244–250 (2004). 100. Smart, M. C. et al. Improved performance of lithium-ion cells with the use of
72. Li, Y.-H. et al. A novel polymer electrolyte with improved high-temperature- fluorinated carbonate-based electrolytes. J. Power Sour.
tolerance up to 170 °C for high-temperature lithium-ion batteries. 119–121, 359–367 (2003).
J. Power Sour. 244, 234–239 (2013). 101. Zhu, G. et al. Materials insights into low-temperature performances of
73. Wu, X.-L. et al. Enhanced working temperature of PEO-based polymer lithium-ion batteries. J. Power Sour. 300, 29–40 (2015).
electrolyte via porous PTFE film as an efficient heat resister. Solid State Ionics 102. Ryou, M.-H. et al. Effect of fluoroethylene carbonate on high temperature
245–246, 1–7 (2013). capacity retention of LiMn2O4/graphite Li-ion cells. Electrochimica Acta
74. Bouchet, R. et al. Single-ion BAB triblock copolymers as highly efficient 55, 2073–2077 (2010).
electrolytes for lithium-metal batteries. Nat. Mater. 12, 452–457 (2013). 103. Kang, K. S. et al. Effect of additives on electrochemical performance of lithium
75. Wang, Y. et al. Design principles for solid-state lithium superionic conductors. nickel cobalt manganese oxide at high temperature. J. Power Sour.
Nat. Mater. 14, 1026–1031 (2015). 253, 48–54 (2014).
76. Li, D., Ma, Z., Xu, J., Li, Y. & Xie, K. High temperature property of all-solid- 104. Liu, Y., Tan, L. & Li, L. Tris(trimethylsilyl) borate as an electrolyte additive to
state thin film lithium battery using LiPON electrolyte. Mater. Lett. improve the cyclability of LiMn2O4 cathode for lithium-ion battery.
134, 237–239 (2014). J. Power Sour. 221, 90–96 (2013).
77. Mo, S. et al. High-temperature performance of all-solid-state battery 105. Rong, H. et al. A novel imidazole-based electrolyte additive for improved
assembled with 95(0.7Li2S-0.3P2S5)-5Li3PO4 glass electrolyte. Solid State Ionics electrochemical performance at elevated temperature of high-voltage
296, 37–41 (2016). LiNi0.5Mn1.5O4 cathodes. J. Power Sour. 329, 586–593 (2016).
78. Kato, Y. et al. High-power all-solid-state batteries using sulfide superionic 106. Pham, H. Q., Hwang, E.-H., Kwon, Y.-G. & Song, S.-W. Understanding the
conductors. Nat. Energy 1, 16030 (2016). interfacial phenomena of a 4.7 V and 55 °C Li-ion battery with Li-rich layered
79. Osada, I., de Vries, H., Scrosati, B. & Passerini, S. Ionic-liquid-based polymer oxide cathode and graphite anode and its correlation to high-energy cycling
electrolytes for battery applications. Angew. Chem. Int. Edn performance. J. Power Sour. 323, 220–230 (2016).
55, 500–513 (2016). 107. Aurbach, D. et al. On the use of vinylene carbonate (VC) as an additive to
80. Del Sesto, R. E. et al. Limited thermal stability of imidazolium and electrolyte solutions for Li-ion batteries. Electrochimica Acta
pyrrolidinium ionic liquids. Thermochimica Acta 491, 118–120 (2009). 47, 1423–1439 (2002).

NATURE ENERGY 2, 17108 (2017) | DOI: 10.1038/nenergy.2017.108 | www.nature.com/natureenergy 13


©
2
0
1
7
M
a
c
im
lla
n
P
u
b
ils
h
re
s
L
im
it
e
d
,
p
a
rt
o
fp
S
ri
n
g
re
N
a
t
u
r
e
.
A
ll
ri
g
h
ts
r
e
s
e
r
v
e
d
.
NATURE ENERGY REVIEW ARTICLE
108. Han, H. et al. Lithium (fluorosulfonyl)(nonafluorobutanesulfonyl)imide 115. Song, M.-S. et al. Is Li4Ti5O12 a solid-electrolyte-interphase-free electrode
(LiFNFSI) as conducting salt to improve the high-temperature resilience of material in Li-ion batteries? Reactivity between the Li4Ti5O12 electrode and
lithium-ion cells. Electrochem. Commun. 13, 265–268 (2011). electrolyte. J. Mater. Chem. A 2, 631–636 (2014).
109. Jin, Z. et al. A new class of phosphates as co-solvents for nonflammable 116. Cui, Y., Rohde, M., Mahmoud, M. M., Ziebert, C. & Seifert, H. J.
lithium ion batteries blectrolytes. ECS Electrochem. Lett. Phosphate based ceramics as solid electrolyte for high temperature lithium
1, A55–A58 (2012). ion batteries. ECS Meeting Abstracts 120, ma2016-01 (2016).
110. MacNeil, D., Lu, Z., Chen, Z. & Dahn, J. R. A comparison of the electrode/
electrolyte reaction at elevated temperatures for various Li-ion battery
cathodes. J. Power Sour. 108, 8–14 (2002). Additional information
111. Plichta, E. J. & Behl, W. K. A low-temperature electrolyte for lithium and Reprints and permissions information is available at www.nature.com/reprints.
lithium-ion batteries. J. Power Sour. 88, 192–196 (2000). Correspondence should be addressed to P. M. A.
112. Yabuuchi, N. & Ohzuku, T. Electrochemical behaviors of
LiCo1/3Ni1/3Mn1/3O2 in lithium batteries at elevated temperatures. How to cite this article: Rodrigues, M.-T. F. et al. A materials perspective on Li-ion
J. Power Sour. 146, 636–639 (2005). batteries at extreme temperatures. Nat. Energy 2, 17108 (2017).
113. Chen, K. et al. Evaluation of the low temperature performance of lithium Publisher’s note: Springer Nature remains neutral with regard to jurisdictional claims in
manganese oxide/lithium titanate lithium-ion batteries for start/stop published maps and institutional affiliations.
applications. J. Power Sour. 278, 411–419 (2015).
114. Luo, Y. et al. Hierarchical carbon decorated Li3V2(PO4)3 as a bicontinuous
cathode with high-rate capability and broad temperature adaptability. Competing interests
Adv. Energy Mater. 4, 1400107 (2017). The authors declare no competing financial interests.

14 NATURE ENERGY 2, 17108 (2017) | DOI: 10.1038/nenergy.2017.108 | www.nature.com/natureenergy


©
2
0
1
7
M
a
c
im
lla
n
P
u
b
ils
h
re
s
L
im
it
e
d
,
p
a
rt
o
fp
S
ri
n
g
re
N
a
t
u
r
e
.
A
ll
ri
g
h
ts
r
e
s
e
r
v
e
d
.

Das könnte Ihnen auch gefallen