You are on page 1of 19

Angewandte

15213773, 2018, 1, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.201703772 by Chonnam National University, Wiley Online Library on [27/12/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews Chemie

International Edition: DOI: 10.1002/anie.201703772


Batteries German Edition: DOI: 10.1002/ange.201703772

From Lithium-Ion to Sodium-Ion Batteries: Advantages,


Challenges, and Surprises
Prasant Kumar Nayak+, Liangtao Yang+, Wolfgang Brehm, and Philipp Adelhelm*

Keywords:
electrode reactions ·
energy storage ·
lithium-ion batteries ·
sodium-ion batteries

Angewandte
Chemie
102 www.angewandte.org T 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2018, 57, 102 – 120
Angewandte

15213773, 2018, 1, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.201703772 by Chonnam National University, Wiley Online Library on [27/12/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews Chemie

Mobile and stationary energy storage by rechargeable batteries is From the Contents
a topic of broad societal and economical relevance. Lithium-ion
1. Introduction 103
battery (LIB) technology is at the forefront of the development, but
a massively growing market will likely put severe pressure on resources 2. Electrode Potential and Cell
and supply chains. Recently, sodium-ion batteries (SIBs) have been Voltage 105
reconsidered with the aim of providing a lower-cost alternative that is
less susceptible to resource and supply risks. On paper, the replacement 3. Materials for Conversion
Reactions 107
of lithium by sodium in a battery seems straightforward at first, but
unpredictable surprises are often found in practice. What happens 4. Materials for the Positive
when replacing lithium by sodium in electrode reactions? This review Electrode 109
provides a state-of-the art overview on the redox behavior of materials
5. Materials for the Negative
when used as electrodes in lithium-ion and sodium-ion batteries,
Electrode 113
respectively. Advantages and challenges related to the use of sodium
instead of lithium are discussed. 6. Summary and Outlook 117

1. Introduction combined with a variety of negative electrodes (Fe, Zn, Cd)


leading to the invention of the Ni-Fe, Ni-Zn, and Ni-Cd
Rechargeable batteries are an indispensable part of our batteries more than 100 years ago, for example. Later on, also
modern society and provide electrical energy on demand in H2 (Ni-H2 battery) and finally metal hydrides (NiMH battery)
a multitude of applications. The ever-growing demand for have been used, the latter being nowadays preferred.
“better batteries” led to a strong increase in R&D within the Practical energy densities of NiMH batteries are around
last 10–20 years with the prime focus on lithium-ion battery 100 Wh kg@1. Lithium-ion batteries are the most recent
(LIB) technology. For many years, portable consumer elec- development and have been commercialized in the early
tronics were the important driver for this development, but 1990s. Thanks to their high energy density, they quickly
more recently two other aspects have put battery research in started to dominate the market for consumer electronics.
the spotlight: 1) Electric vehicles (EVs) that, after several Their cell chemistry is based on the insertion of lithium
attempts during the last century, will finally reach the mass ions into a variety of host structures. Graphite is mainly
market within the coming years; and 2) stationary grid used as negative electrode whereas layered oxides
storage being capable of storing excess electrical energy (Li[NixCoyMnz]O2), LiFePO4, and LiMn2O4 are commonly
supplied by wind and solar power on a very large scale at low used positive electrode materials. A variety of structurally
cost. It is also foreseeable that other markets such a small related compounds as well as alternative electrolytes are
smart devices using thin film battery technology as well as currently studied to further advance this technology.
robotics will soon gain more relevance. The characteristics of Overall, the LIB technology is a real success story. Since
these applications are quite different and no single type of its commercialization, the energy density has continuously
battery fulfills all requirements at once. Different battery increased by about 7–8 Wh kg@1 per year, now reaching about
chemistries are therefore currently studied that, in a simplified 250 Wh kg@1 on the cell level (18650-type cells). At the same
way, aim at increasing energy density (Wh kg@1, Wh L@1) or at time, cell costs decreased much faster than expected from
decreasing costs (EUR/kWh) compared to state-of-the art around 1000 E kWh@1 (mid 1990s) to below 200 E kWh@1
LIB technology. Of course a range of other parameters, today. A further decrease to below 100 E kWh@1 is expected
including safety, cycle and calendar life, temperature window, within the next 5–10 years.[1] Not surprisingly, also EV battery
and so on, are relevant too. packs became drastically cheaper within just a few years.[2]
It is important to realize that despite great efforts in
battery research, only a few rechargeable battery systems
have achieved relevant market shares. From a historic [*] Dr. P. K. Nayak,[+] L. Yang,[+] W. Brehm, Prof. P. Adelhelm
perspective, the lead acid battery technology is the oldest Friedrich-Schiller-University Jena, Institute for Technical Chemistry
and Environmental Chemistry
one and it still dominates the market with respect to the
Philosophenweg 7a, 07743 Jena (Germany)
yearly output in energy storage capacity. This system is more E-mail: philipp.adelhelm@uni-jena.de
than 150 years old, cost-effective, and largely used in auto-
Dr. P. K. Nayak,[+] L. Yang,[+] W. Brehm, Prof. P. Adelhelm
motive applications (starting battery) and for auxiliary power Center for Energy and Environmental Chemistry Jena, CEEC Jena
supply. Major drawbacks of this technology are its limited Philosophenweg 7a, 07743 Jena (Germany)
energy density of about 30 to 40 Wh kg@1 and the low energy [+] These authors contributed equally to this work.
efficiency. Rechargeable alkaline batteries based on NiOOH/ The ORCID identification number(s) for the author(s) of this article
Ni(OH)2 as positive electrode are frequently used too, can be found under:
although to a minor degree. This electrode has been https://doi.org/10.1002/anie.201703772.

Angew. Chem. Int. Ed. 2018, 57, 102 – 120 T 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim 103
Angewandte

15213773, 2018, 1, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.201703772 by Chonnam National University, Wiley Online Library on [27/12/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews Chemie

The overall market share of LIBs is quickly rising and the materials will be required to meet the demand and supply
demand will sharply increase once mass production of electric chains will be challenged. Creating new options for future
vehicles becomes reality. The historic development of LIB energy storage is therefore of strategic relevance. Wadia et al.
technology has been recently summarized by Blomgren.[3] discussed potential resource constraints when scaling up
Despite this promising future, of course there are also battery production for EV and grid storage.[4] Figure 1 shows
boundaries. The increase in LIB energy density, for example, how much battery capacity (energy storage capacity, ESP) can
is expected to reach its intrinsic limits within the next years be supplied from different cell chemistries, assuming that all
with values slightly exceeding 300 Wh kg@1,[1] so research on resources (reserve base) are used to produce batteries only.
alternative cell chemistries and concepts is needed if the past All other uses of the elements are neglected. The element
improvements are supposed to continue. Regardless of any limiting the cell chemistry by resource constraint is shown in
performance parameters, another (often controversial) dis- brackets. The vertical lines indicate the demands for 1 billion
cussion addresses resource issues related to the abundance EV (40 kWh each) and the world daily electricity generation.
and geographical distribution of lithium and some other LIB Although values for the reserve base change with time, the
components, especially cobalt. If electric mobility and elec- graph shows that on the long term LIB technology might run
trical grid storage reach the mass market, a large amount of into resource constraints (limited by either Co or Li) for
a scenario with deep market penetration of stationary storage
and electric mobility. At the same time, it should be noted that
there are large discrepancies in published numbers on
resources and reserves.[121] This might explain the frequent
dispute on potential resource limitations by lithium, for
example. Anyway, it is clear that some elements are not
critical, among Fe, Mn, S, C and Na. Developing batteries
based on these elements is therefore very attractive.
The dimension of materials demand for electric cars
becomes also apparent considering the following. Assuming
conventional LIB technology, a 60 kWh EV battery (about
350 km driving range) contains roughly about 7.5 kg of Li,
65 kg of Ni, Co, Mn in total, 55 kg of graphite, 48 kg copper
and 30 kg of aluminum. Sooner or later, recycling will
therefore become an important factor.
Sodium is an obvious substitute for lithium thanks to its
similar chemical properties. It is worthwhile to note that in the
early days of LIB research, also sodium-ion batteries (SIBs)
Prof. Philipp Adelhelm studied Materials Science at the University of had been studied.[5] The research, however, was largely
Stuttgart and the Max-Planck-Institute for Metals Research. He worked at discontinued, most likely because progress was slow and the
the Max-Planck-Institute for Colloids and Interfaces (Antonietti group, LIB chemistry successful.[6] Today, research on sodium-ion
Ph.D. thesis), the Debye Institute for NanoMaterials Science in Utrecht, batteries is mainly motivated by the large abundance of
(de Jongh group), and the University of Giessen (Janek group). In 2015, he
sodium and the hope to produce batteries that are cheaper
was appointed full professor at the Friedrich-Schiller-University Jena,
Germany. His research interests are on fundamental aspects of energy compared to LIBs and less prone to resource issues.
storage materials with the current focus on alternative battery concepts. Companies such as Faradion Ltd. or the SIB prototype
presented by the French RS2E network are encouraging, and
Dr. Prasant Kumar Nayak completed his Ph.D. from Indian Institute of researchers worldwide are seeking for suitable materials. SIBs
Science, Bangalore, India in 2012. After completing about 4 years of might be also attractive in view of environmental aspects as
postdoctoral research in the Aurbach group at Bar-Ilan University, Israel,
shown by Life cycle assessment.[7] The progress in material
he joined the Adelhelm group in August 2016. Currently, his work addresses
high capacity cathode materials for Li and Na-ion batteries based on
development for SIBs is frequently summarized in review
layered transition-metal oxides. articles.[8] The materials choice is usually inspired or even
identical to what is used in LIBs. The transition from LIBs to
Liangtao Yang received his Master degree at University of Science and SIBs is sketched in Figure 1 and appears seemingly simple.
Technology of China (USTC) in material engineering in 2016. He finished This is, however, not the case and given host structures will
his master thesis in Ningbo Institute of Materials Technology and Engineer- interact very differently depending on whether lithium or
ing (NIMTE), Chinese Academy of Sciences where he studied cathode
sodium is intercalated. The Na ion (r = 1.02 c, CN = 6) is
materials for lithium ion batteries. Recently, he joined the Adelhelm group
as Ph.D. student studying cathode materials for sodium-ion batteries. larger than the Li ion (r = 0.59 c, CN = 4)[9] and less polar-
izing so the phase behavior (coordination, lattice constants,
Wolfgang Brehm received his M.Sc. at Friedrich-Schiller-University Jena in crystal structure) and diffusion properties are highly affected.
materials science in 2015. He finished his M.Sc. thesis in the Otto-Schott- Of course processes at the electrode–electrolyte interface
Institute of Materials Research, investigating thin film nanostructures of (charge transfer, desolvation/solvation) will change too. It has
diblock copolymers. His current research interest is on conversion electrodes been calculated that the desolvation energy for Na ions in
and alloying anode materials for Li-ion and Na-ion batteries as well as thin
a variety of organic solvents is roughly 30 % smaller as
film electrodes under supervision of Prof. Adelhelm.
compared to lithium, for example.[10] The charge transfer

104 www.angewandte.org T 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2018, 57, 102 – 120
Angewandte

15213773, 2018, 1, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.201703772 by Chonnam National University, Wiley Online Library on [27/12/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews Chemie

Figure 1. a) Energy storage potential in TWh for different battery chemistries. The limiting element by resource constraints is shown in brackets.
The star indicates that the ESP value is beyond the limit of the Figure. Figure reprinted from Ref. [4] with permission from Elsevier. b) Sketch of
the well-known working principle of a lithium-ion battery (LIB) with lithium ions intercalating into two host structures, here graphite and LiCoO2
as example. A sodium-ion battery (SIB) can work the same way but the increase in ion size leads to significant changes in the cell behavior.

resistance should therefore be smaller in case of sodium which information on the storage capacity, phase behavior, Gibbs
might enhance electrode kinetics. energy, and electrode kinetics.
An intriguing example from the materials side is graphite.
Although graphite easily forms graphite intercalation com-
pounds with lithium (or with other metals such as K, Rb, Cs), 2. Electrode Potential and Cell Voltage
sodium does not. Aluminum, too, only forms an alloy with
lithium but not with sodium. The copper current collector in A common finding for SIBs is that their cell voltage is
LIBs might therefore be replaced by cheaper aluminum in smaller compared to analogue LIBs. The frequent explan-
SIBs. We will see, however, that in many cases, the replace- ation is that the standard electrode potential of lithium
ment of lithium by sodium leads to more rather than less (@3.05 V vs. NHE) is 340 mV more negative than that of
compounds complicating the redox chemistry and the phase sodium (@2.71 V vs. NHE). It is important to realize that this
behavior of electrodes. Another interesting example demon- argument is not correct and only indirectly linked to the cell
strating the effect of the ion size has been reported by voltage of LIBs and NIBs, respectively. Explanations for this
Komaba et al.[11] LiCrO2 and NaCrO2 both possess quite can be found in textbooks.[13] Let us first consider Reac-
similar crystal structures, but the former is inactive in Li cells tion (1)
while the latter is active in Na cells. We note that replacement
of lithium by sodium is also an attractive strategy for the AðsÞ Ð Aþ ðgÞ þ e@ ð1Þ
“beyond lithium-ion” systems, such as Li-air and Li-sulfur.
The cell chemistry of these systems is based on very different with A being an alkali metal. The sublimation energy (or
reaction mechanisms and is therefore not discussed here. A more precisely cohesion energy) and ionization energy of A
comprehensive comparison on this topic can be found else- determine the thermodynamics of this reaction. Considering
where.[12] trends in the periodic table, sublimation and ionization
As mentioned above, reviews on the progress in materials energies decrease within the series of alkali metals. Values
development for SIBs are regularly published. The intention for the cohesive energies and ionization energies are higher
of this review is different. Rather than summarizing all for lithium (152.7 kJ mol@1, 513.3 kJ mol@1) than for sodium
materials studied, we focus on the comparison of materials (108.8 kJ mol@1, 495.8 kJ mol@1), for example. The total differ-
when used as electrodes in LIBs and SIBs. We focus on ence is 61.4 kJ mol@1. This argument should place lithium at
selected examples among graphite, LiCoO2 vs. NaCoO2 and the least negative position within the alkali metals in the
their related NMC compounds, LiFePO4 vs. NaFePO4, electrochemical series. Another trend, however, is observed
Li2Ti3O7 vs. Na2Ti3O7, alloys and conversion reactions. The and the redox potentials of the alkali metals show the
impact of the alkali ion on the electrode reaction can be sequence Li (@3.05 V vs. NHE), Cs (@2.92 V), Rb (@2.93 V),
conveniently seen from charge–discharge curves (sometimes K (@2.93 V) to Na (@2.71 V). (Interestingly, the expected
called voltage or potential profiles), which provide direct trend is observed for the alkaline earth metals, with Ba

Angew. Chem. Int. Ed. 2018, 57, 102 – 120 T 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.angewandte.org 105
Angewandte

15213773, 2018, 1, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.201703772 by Chonnam National University, Wiley Online Library on [27/12/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews Chemie

(@2.91 V vs. NHE), Sr (@2.89 V), Ca (@2.87 V), Mg LiðsÞ þ 6 CðsÞ ! LiC6 ðsÞ ð6Þ
(@2.36 V) and Be (@1.97 V.) The reason for the trend in the
alkali metals is simply that half-cell potentials include the and another one serving as positive electrode, for which the
solvation energy of the ion. Although often omitted, Reaction reaction with lithium metal releases a large amount of energy
(1) should read as given in Reaction (2) (redox potential much more positive than Li/Li+ = 0 V), for
example, intercalation of Li+ into layered Li0.5CoO2 close to
AðsÞ þ n ðsolvÞ Ð Aþ ðsolvÞn þ e@ ð2Þ 4 V vs. Li/Li+ in Reaction (7):

LiðsÞ þ 2 Li0:5 CoO2 ðsÞ ! 2 LiCoO2 ðsÞ ð7Þ


where solv represents a solvent molecule and n is the
solvation number of the ion. The standard electrochemical
The same arguments hold for sodium and any other inter-
series is based on aqueous solutions with, in case of dissolved
calation system. What is then left in determining the cell
species, concentrations of the relevant ions corresponding to
voltage is simply the Gibbs energy change of the cell reaction
unit activity at 298 K. Solvation energies commonly increase
DrGo, which is linked to the cell voltage E by the well-known
with decreasing ion radius, so the hydration energy of Li+
Reaction (8):
(@481 kJ mol@1) is much larger than for Na+ (@375 kJ mol@1).
It is this large difference of more than 100 kJ mol@1 that shifts Dr Go
lithium to the most negative position among the alkali metals E¼@ ð8Þ
z?F
in the electrochemical series. It is also clear from Reaction (2)
that the type of solvent affects the solvation energy, and hence with z being the number of transferred electrons and F the
the redox potential. For every solvent there will be an own Faraday constant. The intercalation potentials of Li+ and Na+
electrochemical series, and redox potentials for non-aqueous are highly structure-dependent, but for given intercalation
solvents will deviate from the usual case of assuming an hosts they are generally found to be lower for sodium than for
aqueous electrolyte. A comparison between water and lithium. This is observed for positive as well as for negative
a range of non-aqueous solvents shows that in most cases electrodes, so whether the cell voltage of an SIB is higher or
the differences are in the range of a few hundred mV or less.[14] lower compared to an analogue LIB depends on the sum of
Although these values are relatively small, it is clear that the differences at both electrodes. Ong et al. calculated that
redox potentials should not be simply transferred from the intercalation potentials for a range of positive electrode
electrochemical series with aqueous solution to non-aqueous materials are lower for sodium than for lithium, that is, less
solutions used in LIBs and SIBs. energy is released when intercalating Na+ instead of Li+ into
Now, in a lithium-ion battery two half-cell reactions are the host structure.[15] A difference of 0.57 V for layered AMO2
combined sharing the same electrolyte. We consider the structures and 0.39 V for the olivine structure were calculated.
popular example of graphite and LiCoO2 in Reactions (3)– The same situation has been found for certain titanates that
(5): are negative electrode materials. The intercalation potential
of Li+ into Li2Ti3O7 was calculated to be 1.46 V vs. Li/Li+,
negative electrode: while sodiation of Na2Ti3O7 occurs at 0.37 V vs. Na/Na+,[16] the
discharge difference is larger than one volt. For alloys, that are also
LiC6 ðsÞ þ n ðsolvÞ KKKK!
KKKK Liþ ðsolvÞn þ 6 CðsÞ þ e@ ð3Þ potential negative electrodes. Chevrier and Ceder calculated
charge
that the redox potential is about 0.15 V lower for sodium than
positive electrode: lithium.[17]
discharge Figure 2 illustrates how the individual electrode potentials
2 ? Li0:5 CoO2 ðsÞ þ Liþ ðsolvÞn þ e@ KKKK!
KKKK 2 LiCoO2 ðsÞ þ n solv ð4Þ
charge and cell voltages change when comparing LIBs with SIBs. The
cell voltage is usually lower for sodium cells because for
The overall cell reaction is (3) + (4):
positive electrode materials, the shift in redox potentials to
discharge lower values is larger than for most negative electrode
LiC6 ðsÞ þ 2 Li0:5 CoO2 ðsÞ KKKK!
KKKK 2 LiCoO2 ðsÞ þ 6 CðsÞ ð5Þ
charge materials. Unfortunately, also graphite cannot be used in
SIBs (unless under special circumstance, see Section 5.1). On
It can be seen from the overall reaction that the solvation the other hand, a cell based on ACoO2 and A2Ti3O7 would
energies are not relevant in determining the cell voltage. This give a higher cell voltage in case of sodium, for example,
is also true for using lithium metal as positive electrode. A because the shift in redox potential between Li and Na in
more quantitative analysis can be also found in Ref. [15]. The ACoO2 (0.57 V) is smaller than the shift in A2Ti3O7 (1.09 V).
above discussion on the influence of the solvation properties Overall, high voltage sodium-ion batteries are not without
on the redox potential therefore becomes irrelevant when reach, in principle. Considering the alkali metals alone,
considering full cells (however kinetics might be affected). sodium should enable cell voltages that are higher by 0.53 V
In other words, what is relevant for obtaining a high cell compared to lithium owing to its smaller cohesive energy.[15]
voltage is to combine two suitable intercalation compounds: This advantage cannot be harvested so far because in almost
One, serving as negative electrode, for which the reaction with all cases, Na+ intercalation into a fixed positive electrode is
lithium metal only releases little energy (redox potential close energetically less favorable as Li+ intercalation. This leads to
Li/Li+ = 0 V), for example, intercalation of Li+ into graphite the commonly observed lower overall cell voltage for SIBs. A
at below 0.25 V vs. Li/Li+ in Reaction (6): few exceptions, however, exist that will be discussed in

106 www.angewandte.org T 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2018, 57, 102 – 120
Angewandte

15213773, 2018, 1, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.201703772 by Chonnam National University, Wiley Online Library on [27/12/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews Chemie

for MgH2). Positive and negative electrodes can therefore be


realized.
Although conversion electrodes show high capacities in
experiments, several unsolved challenges hinder their use in
commercial, rechargeable cells. Of course, every conversion
compound has its individual characteristics, but generally
conversion reactions suffer from a low initial coulombic
efficiency (typically below 75 %), poor kinetics (overpoten-
tials increase with increasing bond polarity; the combined
overpotentials for fluorides can exceed 1 V, for example) and
side reactions with the electrolyte. The formation of new
phases also imposes large changes in electrode volume. The
low initial coulombic efficiency is linked to an activation of
Figure 2. Selection of different lithiation/sodiation potentials for a vari- the electrode that is due to severe structural rearrangements.
ety of compounds. Data has been collected from different literature
The first lithiation leads to the formation of a nanocomposite
sources.[15–17, 19] Note that the redox potentials of Li/Li+ and Na/Na+
are not identical but are arbitrarily defined as 0 V on each potential consisting of an often amorphous LicX matrix in which the
axis. metal nanoparticles are distributed. At the same time, a gel-
like inorganic/organic surface film forms due to electrolyte
Section 3. It should be also noted that electrode materials decomposition. Depending on the compound, a number of
such as Na4Co3(PO4)2P2O7 pyrophosphate with redox activity intermediate phases can also form during the electrode
between 4.1 and 4.7 V (vs. Na/Na+) have been reported.[18] reaction. Therefore, the seemingly simple Reaction (9) is
Although challenges remain, this demonstrates that Na+ often much more complex, difficult to characterize structur-
intercalation into positive electrodes can also lead to very ally, and the experimentally determined electrode potentials
high redox potentials. deviate significantly from the theoretical ones (copper sulfide
is a notable exception).[24] Because the ionic/electronic
conductivity of the nanocomposite is small, electrodes are
3. Materials for Conversion Reactions typically prepared with larger amounts of conductive additive.
The typical voltage profile of a conversion reaction is
A popular approach for realizing electrode materials with sketched in Figure 3 showing the large initial irreversible
very high capacity is to use so-called conversion reactions of capacity and the pronounced voltage hysteresis.
the type shown in Reaction (9): Klein et al. discussed the influence of the alkali metal on
conversion reactions.[23] Considering the abovementioned
Ma Xb þ ðb ? cÞA Ð a ? M þ b ? Ac X ðA ¼ Li or NaÞ ð9Þ reactions, a comparison of the cell voltages between lithium
and sodium is straightforward. Writing Reaction (9) for both
with M being a transition metal (or Mg) and X a non-metal cases and considering their difference, it can be easily seen
(such as F, O, P, N, S, H). In general, conversion reactions that the shift in cell voltage when replacing lithium by sodium
show much higher capacity compared to intercalation reac- is constant for hydrides, oxides, fluorides, and so on. For
tions because the transition metal is fully reduced to the oxides, the shift in cell voltage is 0.96 V, which formally
metallic state. The reaction of CoO with lithium to form Co corresponds to Reaction (10):
and Li2O theoretically corresponds to a capacity of
706 mAh g@1. In comparison, reducing Co4+ to Co3+ in the 2 Li þ Na2 O Ð Li2 O þ 2 ? Na DEðLi@NaÞ ¼ 0:96 V ð10Þ
classical LiCoO2 intercalation compound corresponds to
274 mAh g@1 (roughly half of this value is reached in practice, In other words, this means that for any conversion reaction of
as only about 0.5 lithium atoms per formula can be reversibly Reaction (9) with an oxide MaOb, the cell voltage in case of
intercalated). Owing to the large number of MaXb compounds sodium is lower than for lithium by 0.96 V. Figure 3 shows
available, this concept provides an enormous potential. a comparison for the different classes of compounds. For
Although the concept of conversion reactions has been sulfides, oxides, and hydrides, the shift is in the range of
already known for a long time, research boomed since the 400 mV, that is, also here, the “lithium version” of the cell
late 1990s when studies on the reversible lithiation of tin- leads to a higher voltage. On the other hand, for chlorides, the
based composite oxides[20] and a series of binary transition- cell voltage for a conversion reaction with lithium and sodium
metal oxides at room temperature were published.[21] Com- is virtually identical. For the even heavier halides, a conversion
prehensive overviews on the materials studied for LIBs can be reaction with sodium theoretically delivers a higher cell
found in Ref. [22]. Assuming bulk thermodynamics, the redox voltage than in case of lithium. This behavior can be
potential can be conveniently calculated from the Gibbs rationalized by considering the relevant Born–Haber
energy of the reaction DrG using Reaction (8). The redox cycles.[23] The lattice energies of the lithium compounds are
potential of conversion reactions increases with the bond larger than the ones for sodium in all cases favoring more
polarity between M and X, so reactions with fluorides occur at negative Gibbs reaction energies and hence larger cell
higher redox potentials (around 3 V vs. Li/Li+) compared to voltages. However, due to the lower cohesive and ionization
sulfides and oxides (around 1.5–2.0 V) and hydrides (0.52 V energies in case of sodium, this difference can be compen-

Angew. Chem. Int. Ed. 2018, 57, 102 – 120 T 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.angewandte.org 107
Angewandte

15213773, 2018, 1, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.201703772 by Chonnam National University, Wiley Online Library on [27/12/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews Chemie

Figure 3. a) Simplified voltage profile of a typical conversion electrode. During the first ion insertion, the electrode undergoes an activation that
involves the formation of a nanocomposite structure and that also leads to a high irreversible capacity Qirr in the first cycle. The nanocomposite is
then fairly reversible and exhibits a larger hysteresis. b) Replacing lithium by sodium in conversion reactions leads to a constant shift in redox
potentials for different types of compounds. For hydrides, oxides, sulfides, and fluorides; a “lithium cell” gives a higher cell voltage. For chlorides,
however, the difference is virtually zero. For even more heavy halides, a hypothetical “sodium cell” would deliver a higher cell voltage. c) Volume
expansion of conversion electrodes for different types of compounds for the reaction with lithium and sodium, respectively. All values are
calculated for 25 8C.[23] d) Specific capacities and redox potentials vs. Na/Na+ for conversion reactions of different classes of materials with
sodium.

sated (chlorides) or even overcompensated (iodides, bro- be + 0.94 V, + 0.93 V, + 0.91 V, + 0.55 V, + 0.47 V, @0.04 V,
mines). This simple examination proofs what has been + 0.52 V. The redox potentials vs. the respective metal
already discussed in Section 2, that is, the replacement of electrode would be therefore much lower for magnesium,
lithium by sodium does not necessarily lead to a lower cell with oxides being an exception.
voltage. The negligible differences between lithium and Lithiation of conversion materials is generally accompa-
sodium cell voltages for chlorides might also reflect the nied by large volume expansions. The effect is smallest for the
quite reasonable cell voltage of the high temperature Na/ halides (typically below 50 %) but can easily amount to more
NiCl2 battery (2.6 V at T = 245 8C). It is worthwhile to than 100 % for sulfides or oxides with the known issues of
mention that the same calculations can be made for con- mechanical electrode deterioration. For sodium, the volume
version reactions with magnesium (magnesium-ion batteries). changes are roughly twice as high as shown in Figure 3. It is
In a direct comparison between lithium and magnesium, therefore foreseeable that designing a durable conversion
values of DE (Li@Mg) for the series iodides!hydrides would electrode for sodium cells will be more challenging. Note-

108 www.angewandte.org T 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2018, 57, 102 – 120
Angewandte

15213773, 2018, 1, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.201703772 by Chonnam National University, Wiley Online Library on [27/12/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews Chemie

worthy, alloys show an even larger volume expansion (see Ong et al. have studied the impact of the ion size on the
Section 5.2). redox potential of a variety of popular electrode materials by
At this point, we remind that conversion electrodes based DFT methods,[15] as shown in Figure 4. As mentioned before,
on Reaction (9) generally only function because the first ion they calculated that the redox potentials for sodium insertion
insertion involves the formation of a peculiar nanostructure. are between 0.18 to 0.57 V lower as compared to lithium,
This way, diffusion distances remain small enough to render which is the main reason for the often observed lower cell
the reaction reversible. The nanostructure is a result of voltage in case of SIBs. In the following, we will discuss
a complex interplay between different properties that relate experimental results on selected examples of positive electro-
not only to volume expansion but also to diffusion coefficients des.
of the separate species that can decide on whether reversi-
bility is achieved or not. The formation of too large nano-
particles during activation might prevent rechargeability, as
shown for CuF2 by Wang et al., for example.[25] It is clear that
replacing lithium by sodium will affect this interplay.
The hope that the challenges of lithium conversion
reactions might be overcome by sodium, however, has not
been fulfilled so far. Only few studies focus selectively on the
comparison of conversion reactions with lithium and sodium.
Overall, capacities are frequently found to be lower for
sodium as compared to the equivalent lithium cell. For
example, nanosized Fe2O3 (qth = 1006 mAh g@1) showed
a capacity of 1000 mAh g@1 in lithium cells, whereas around
350 mAh g@1 were found in sodium cells.[26] Although this
value is lower, it is still much higher compared to intercalation
electrodes. For CuO (qth = 674 mAh g@1), capacities for lith-
ium where roughly twice as high as for sodium
(> 600 mAh g@1 vs. 350 mAh g@1). The reaction therefore is
incomplete for sodium and largely restricted to the Cu2O
intermediate phase. The volume expansion for the lithiation
of CuO to form Cu and Li2O (+ 74 %) is very close to the
expansion for sodiation of CuO to form Cu2O and Na2O
(+ 74 %). This indicates that the increased volume expansion Figure 4. Calculated Na voltage vs. Li voltage for different structures.
Reproduced from Ref. [15] with permission of The Royal Society of
in sodium conversion reactions indeed impedes full utiliza-
Chemistry.
tion. At higher current densities (1350 mA g@1), however, the
difference diminishes and capacities for lithium and sodium
converge to values of around 250–300 mAh g@1[27] . Interest- There are mainly two types of positive electrode materials
ingly, conversion reactions with the unconventional carbodii- such as layered transition-metal oxides and polyanionic
mide anion can show a different behavior, that is, the compounds. Layered transition-metal oxides have attracted
capacities for sodium can exceed the ones of lithium.[28] great interest because of their high theoretical capacity close
Despite the existing challenges, conversion reactions of the to 250 mAh g@1. In general, the layered oxides containing Li
described type still remain an attractive research objective, have a-NaFeO2 structure formed by edge sharing MO6
especially considering low cost compounds like iron oxides or octahedra separated by Li ions occupying interstitial octahe-
sulfides such as FeS2.[29] dral sites. On the other hand, the Na-based layered transition-
metal oxides are classified as O and P type depending on the
stacking of close-packed oxygen layers. Among them, the
4. Materials for the Positive Electrode most common types are O3 and P2. The letters O and P
denote the octahedral and prismatic sites occupied by Na
Materials for the positive electrode are mainly com- while the numbers 2 and 3 denote the number of oxygen
pounds containing the 3d transition metal cations, namely Co, stacking layers, respectively. For instance, the oxygen stacking
Mn, Fe, or Ni as redox-active elements. Layered oxides such of O3, P2, and P3-type structures are ABCABC, ABBA, and
as LiCoO2 and Li[Ni1-x-yMnxCoy]O2, LiFePO4, or spinel ABBCCA, respectively. In the O3 phase, Na+ and transition
LiMn2O4 with capacity values in the range of 120– metals are located in the octahedral sites of alternating Na
180 mAh g@1 are currently implemented in commercial LIBs. and transition-metal oxide layers, respectively (Figure 5 a),
A range of other materials is being investigated with the main whereas Na+ resides in the trigonal prismatic sites in P2 phase
aim to increase the redox potential and/or capacity. In view of (Figure 5 b). The advantage with Na-based layered oxides is
element abundance, it is clear that the use of Fe- and Mn-rich that these can be synthesized from a wide range of transition
compounds are the preferred choice for SIBs. Nevertheless, metals varying from Ti to Cu, while synthesis is limited to Mn,
a direct comparison between for example, LiCoO2 and Ni, and Co in case of their Li analogous.[31] A popular
NaCoO2 is very useful to discuss structural effects. approach for tuning the electrochemical properties of these

Angew. Chem. Int. Ed. 2018, 57, 102 – 120 T 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.angewandte.org 109
Angewandte

15213773, 2018, 1, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.201703772 by Chonnam National University, Wiley Online Library on [27/12/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews Chemie

Figure 5. Representation of layered oxides: a) O3, b) P2, and c) O2 type. The letters (A, B, C) are used to describe the packing patterns of the
oxygen-ion frameworks. Reproduced from Ref. [30] with permission from Wiley.

materials is to mix different transition metals. A large variety properties of pyrophosphate based materials for both Li and
of materials has therefore been studied that have been Na-ion batteries.[34] Figure 6 shows a comparison of the
summarized in a number of comprehensive reviews.[8g, 30, 32] In average voltage and specific capacity of some polyanionic
the following, we therefore discuss only some selected layered compounds for both Li- and Na-ion batteries. Generally, the
oxides and put more emphasis on the comparison between Li-based polyanionic compounds have higher average voltage
their use in lithium and sodium cells. and also capacity as compared to their Na counterparts.
Apart from layered transition metal oxides, there is also LiFePO4 has a higher voltage (3.5 V) and capacity of
increasing interest in polyanionic compounds because of their 160 mAh g@1 as compared to olivine NaFePO4 having
structural diversity, high working potential, and high reversi- (3.0 V) and 110 mAh g@1, for example.
bility for intercalation/deintercalation of Li and/or Na ions. Although not in the focus of this review, it is worth
Polyanionic compounds became popular after the pioneering mentioning that surface films can be formed on the positive
work of Goodenough et al. on the olivine-type cathode electrode owing to side reactions with the electrolyte solution.
material LiFePO4 that shows a redox plateau of about 3.5 V The so-called cathode-electrolyte interface for a variety of
vs. Li/Li+.[33] Barpanda et al. reviewed the electrochemical lithium positive electrodes has been discussed by Aurbach
et al.[35] and Edstrom et al.[36] already some time ago, for
example. Understanding this film formation as well as
tailoring its properties by additives is still challenging but
very important for further improving the long-term stability
of positive electrodes. This becomes especially relevant for
high voltage materials such as LiNi0.5Mn1.5O4 for which
lithium bis(oxalato) borate (LiBOB) has been found to be
very effective to reduce ageing, for example.[37] The overall
knowledge about surface films on lithium positive electrodes,
however, is still poor and currently largely lacking for the
analogue sodium compounds.

4.1. Comparison for Selected Examples


4.1.1. Layered Oxides

Although LiCoO2 and NaCoO2 have the same frame-


works of CoO6 edge-sharing octahedra, they show different
redox potentials and experience a different phase evolution
during cycling (Figure 7). Depending on the degree of
intercalation, the potential difference between both com-
pounds varies within 0.4 to 1.0 V. The complex voltage profile
of NaCoO2 indicates that numerous phase transitions and
Figure 6. Comparison of average potential and discharge capacity for ordering phenomena occur during cycling. Much research had
lithium (LiFePO4, LiFePO4F, Li2Fe(SO4)2, LiFeSO4F, LiFeSiO4, shown that LiCoO2 maintains O3 structure until the Li
Li2CoP2O7, LiVOPO4, Li3V2(PO4)3) and sodium compounds (NaFePO4,
content is lower to 0.5. Similarly, when the Na content is lower
NaFePO4F, NaFe(SO4)2, NaFeSO4F, Na2FeSiO4, NaFeP2O7, NaVOPO4,
Na3V2(PO4)3). The contour indicates the theoretical energy density than 0.5, the structure transfers from O3 to P3, which is
obtainable (assuming a Li or Na metal counter electrode). Data undesired for achieving reversibility. As a result, only half of
derived from Refs. [33, 34, 38–48], respectively. theoretical capacity can be reached. During discharging, the

110 www.angewandte.org T 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2018, 57, 102 – 120
Angewandte

15213773, 2018, 1, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.201703772 by Chonnam National University, Wiley Online Library on [27/12/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews Chemie

vs. Na/Na+[55] while its analogous LiNi0.33Mn0.33Co0.33O2 is


known to exhibit a capacity of about 160 mAh g@1 in the
potential range of 2.8–4.4 V.[56]
Co-free layered oxide cathodes are important because of
high cost, low abundance, and toxicity of Co. Sun et al. have
studied Co-free Ni rich layered oxide cathodes for Li and Na
ion batteries.[57] Aurbach et al. synthesized both
LiNi0.5Mn0.5O2 and NaNi0.5Mn0.5O2 by the co-precipitation
method followed by mixing with LiOH or NaOH and
calcining at high temperatures, and they studied their electro-
chemical performances in half cells using Li or Na and full
cells using hard carbon as the negative electrode.[58] The
charge–discharge curve for the 2nd cycle is shown in Figure 8.
As expected, LiNi0.5Mn0.5O2 exhibited a high specific capacity
of about 170 mAh g@1 in the potential range of 2.7–4.3 V,
whereas NaNi0.5Mn0.5O2 exhibited about 136 mAh g@1 in the
Figure 7. Voltage profiles of Li/LiCoO2 and Na/NaCoO2 cells. The
potential range of 2.0–4.0 V. The operating voltage in case of
green line is LiCoO2 from 3.0–4.5 V vs. Li+/Li and the blue line (data lithium is also much higher.
derived from Ref. [49]) is NaCoO2 from 2.0–3.8 V vs. Na+/Na.

compounds Na1/2CoO2 (3.45 V vs. Na+/Na), Na4/7CoO2


(3.15 V vs. Na+/Na), Na2/3CoO2 (2.8 V vs. Na+/Na),
Na0.72CoO2 (2.56 V vs. Na+/Na), Na0.76CoO2 (2.47 V vs. Na+/
Na), Na0.79CoO2 (2.38 V vs. Na+/Na), are formed. The phase
behavior of NaCoO2 has been thoroughly studied by Delmas
et al.[49] With respect to application, the redox behavior of
LiCoO2 as shown in Figure 7 is very well suited as the
potential changes only little during cycling. A very sloping and
gradual phase transforming behavior as found for NaCoO2 is
impractical but the behavior can be tuned by partial
substitution of Co with Mn and Fe, for example.[50] It is also
worth mentioning that ion diffusion in layered NaCoO2 is
calculated to be easier as compared to LiCoO2.[15]
Another layered oxide, NaFeO2, can deliver a specific
capacity of 80–100 mAh g@1 when the cut-off voltage is limited
to 3.4 V, but it loses capacity upon cycling to potentials above
3.5 V owing to irreversible structural change.[50a, 51] Similarly, Figure 8. Voltage profiles (2nd cycle) of LiNi0.5Mn0.5O2 and
its analogous LiFeO2 synthesized from FeOOH exhibited an NaNi0.5Mn0.5O2. Data obtained from Ref. [58].
initial capacity of about 100 mAh g@1 and suffered from large
capacity fading upon cycling due to cationic disorder in the
voltage region of 4.2–1.5 V.[52] About 150 mAh g@1 was found 4.1.2. Layered Transition-Metal Oxides with Excess Lithium in the
for nanocrystalline LiFeO2 prepared at low temperature Transition-Metal Layer (Li1+x[NiMnCo]1-xO2 and
(150 8C), but also here, the material underwent capacity Na[Lix{NiMnCo}1-x]O2)
fading due to structural change to a spinel phase (LiFe5O8)
upon charge–discharge.[53] The concept of excess lithium-based layered oxide
Improvements can be obtained by partial substitution of cathodes came from Li2MnO3 (or Li[Li1/3Mn2/3]O2), where
Fe as shown by Yamada et al., for example.[54] Starting from one third of Mn is substituted by Li in the transition-metal
O3-NaFeO2 they used nickel to substitute iron to prepare layer. This was considered as a promising cathode material
NaFe0.3Ni0.7O2, which exhibited a specific capacity of about because of its high theoretical specific capacity above
140 mAh g@1 when cycled in the voltage range of 2.0–3.8 V vs. 450 mAh g@1. However, practically a specific capacity of
Na/Na+. On partial substitution of Fe with Co, NaFe0.5Co0.5O2 about 250 mAh g@1 was achieved for nanosize Li2MnO3,
exhibited a reversible capacity of about 160 mAh g@1 in the whereas the capacity decreased upon increasing the particle
voltage range of 2.0–4.5 V.[50a] Apart from the specific size.[59] The capacity of the nanosized particles of Li2MnO3
capacity, there is also improvement in their voltage profile was also not stable because of structural transformation upon
upon substitution of Fe with Ni and Co. Sathiya et al. cycling. However, integrated Li1+xM1@xO2 (M = Mn, Ni, Co,
synthesized NaNi0.33Mn0.33Co0.33O2, which exhibited a reversi- Fe) materials composed of Li2MnO3 and LiMO2 in nano-
ble capacity of 120 mA h g@1 corresponding to the utilization domains exhibited specific capacities + 250 mAh g@1 upon
of about 0.5 Na when cycled in the voltage range of 2–3.75 V cycling with potentials higher than 4.5 V.[47, 60] This value of

Angew. Chem. Int. Ed. 2018, 57, 102 – 120 T 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.angewandte.org 111
Angewandte

15213773, 2018, 1, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.201703772 by Chonnam National University, Wiley Online Library on [27/12/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews Chemie

capacity is indeed higher than the capacity of commercialized cycled at 25 mA g@1 in the potential range of 1.5–4.5 V vs.
@1 @1
LiCoO2 (140 mAh g ), LiMn2O4 (120 mAh g ) and LiFePO4 Na.[64b] Recently, de la Llave et al. reported the increase in
@1 [61]
(160 mAh g ) and so on. Hence, from the last decade, there energy density of Mn-based layered oxide cathodes by Li
is an increasing interest in these high capacity cathode doping for Na-ion batteries.[66] Na0.6MnO2 exhibited an initial
materials with the main aims of stabilizing their capacity capacity around 162 mAh g@1, which decreased to a value of
and average discharge voltage upon prolonged cycling, 120 mAh g@1, thus retaining about 75 % capacity after 100
improving their rate capability, and increasing the Coulombic cycles. On the other hand, Na0.6Li0.2Mn0.8O2 exhibited a stable
efficiency in the first cycle. Doping of several cations, such as specific capacity of about 190 mAh g@1 even after 100 cycles
[62]
Na, K, Mg, Al, Sn, is found to be effective in enhancing the when cycled at 20 mA g@1 (C/10) rate. Liu et al. recently
stabilization in their capacity as well as average discharge reported Li-substituted P2/O3 biphasic
voltage by suppressing the structural layered-to-spinel trans- Na0.67Mn0.55Ni0.25Li0.2O2, which exhibited a capacity of about
formation; however, it could not be prevented completely. 158 mAh g@1 at 12 mA g@1, as compared to 147 mAh g@1 for
Also, surface coating by several inert materials, such as Al2O3, a Li-free cathode. Li substitution resulted in more defects to
ZnO, SnO2, AlF3, Li3PO4 and AlPO4, helped in improving the maintain charge neutrality, which enhanced the electronic
stability in cycle life of these Li and Mn-rich cathodes.[63] conductivity and also Na ion diffusion.[45] Considering the
In a similar fashion to that of Li and Mn-rich cathodes for working voltage of layered oxides based on Li in the TM
LIBs, the presence of Li ions in the transition-metal layer of layers, Li and Mn-rich cathodes have an average discharge
cathodes for SIBs results in an increase in specific capacity as voltage of 3.5 V in Li-ion half cells.[62e] On the other hand,
[64]
well as stability. The presence of Li-ions in the TM layers oxides containing Li in TM layers possess an average
allows more Na-ions to remain in the compound when discharge voltage of about 3.1 V in Na half cells.[67] Thus,
charged to very high voltage and hence the P2 structure is a voltage difference of about 0.4 V is observed in these
retained upon cycling. Johnson et al. reported a stable layered transition-metal oxide-based cathodes containing Li
capacity of about 100 mAh g@1 for Na0.85Li0.17Ni0.21Mn0.64O2 in the metal layer.
when cycled at 15 mA g@1 in the potential range of 2.0–4.2 V
vs. Na.[65] Within this restricted voltage window, they found 4.1.3. Structural Change upon Cycling of Layered Transition-
that less than 5 % of the lithium was removed on full Metal Oxides
oxidation and the presence of remaining Li in the transition-
metal layer stabilizes the structure. Meng et al. have exten- Figure 9 shows the phase evolution of some commonly
sively studied the improved electrochemical performance of used layered lithiated transition-metal oxides and their Na
Li doped layered metal oxides such as P2- counterpart upon initial cycling. It can be clearly seen that the
Na0.8[Li0.12Ni0.22Mn0.66]O2 for Na-ion batteries, which exhib- O3-type layered oxides containing Li (such as LiCoO2) can
ited a specific capacity of about
120 mAh g@1 when cycled in the
potential range of 2.0–4.4 V. They
confirmed by NMR studies that
most of Li (about 85 %) presented
in the transition-metal layer.[64a]
The Li presented in the transi-
tion-metal layer undergoes de-
intercalation at high voltage
(above 4.2 V), which can be seen
as a plateau during charge in the
first cycle, similar to that observed
in Li and Mn-rich cathodes. The
presence of Li helps in retaining
more Na ions in the prismatic sites,
which helps in stabilizing the P2
structure even after cycling to high
voltages. On the other hand, lay-
ered oxide cathodes with the P2
structure and without Li undergo
structural transformation to O2
upon cycling to higher voltage.
They also reported Li-substituted
O3-structured layered oxides
Figure 9. Phase behavior of different lithium layered oxides (LiCoO2, LiNi0.5Mn0.5O2,
NaLi0.07Ni0.26Mn0.4Co0.26O2, which
LiNi0.33Mn0.33Co0.33O2) and sodium layered oxides. Mainly P2-type materials (Na2/3MnO2, Na2/3-
exhibited an initial capacity of Ni1/3Mn2/3O2, Na2/3Mn1/2Fe1/2O2) are shown on top, mainly O3-type materials (NaCoO2,
147 mAh g@1 and retained NaNi0.5Mn0.5O2, NaNi0.33Mn0.33Co0.33O2) are shown at the bottom. Data for these materials are derived
128 mAh g@1 after 50 cycles when from Refs. [55, 69, 71–74].

112 www.angewandte.org T 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2018, 57, 102 – 120
Angewandte

15213773, 2018, 1, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.201703772 by Chonnam National University, Wiley Online Library on [27/12/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews Chemie

maintain their structure relatively long upon charging. In case


of LiNi0.5Mn0.5O2 and LiNi0.33Mn0.33Co0.33O2, the structural
changes occur until the Li concentration is lower than 0.6 and
0.2, for example. There is a strong driving force for Li based
layered oxides to undergo a structural layered-to-spinel
transformation upon cycling to voltages + 4.3 V, whereas
there is no thermodynamic force for such transformation in
case of their Na analogues.[68] Under high potential, it
transforms to O3’ which has the similar interlayer distance
of O3 structure. Briefly, doping with other transition metals
can suppress O3 structure transition. In the case of Na-based
O3 layered oxides such as NaCoO2, only small amount of Na
can be removed from the structure before it becomes
unstable, hence only a smaller capacity of about
115 mAh g@1 is achieved.[69] The Na based P2-type layered
oxides such as Na2/3MnO2 can maintain its structure up to the
extraction of 0.4 Na and beyond this, its P2-structure can be
transformed to OP4 phase.[70] Also, P2-Na2/3Ni1/3Mn2/3O2
transformed to O3 phase upon extraction of 0.34 Na. Gen-
erally, P-O structural transformation results in worsening of
electrochemical performance of Na based layered oxides.

4.1.4. LiMn2O4 and NaMn2O4

Mn-based spinel LiMn2O4 is an attractive cathode mate-


rial because of the redox activity of Mn around 4.0 V vs. Li.[75]
It exhibits a specific capacity of about 110 mAh g@1 against its
theoretical capacity of about 140 mAh g@1. Unfortunately, the
Na counterpart of it, namely NaMn2O4, is not stable and
transforms into a layered structure after few cycles.[76] More
recently, Liu et al. synthesized NaMn2O4 under high pressure,
which exhibited a capacity of only about 65 mAh g@1 at
Figure 10. a) Comparison of charge and discharge profile for olivine
a current of 5 mA g@1 in the voltage range of 2.0–4.0 V.[77]
NaFePO4 and LiFePO4 cathode materials at 10 mA g@1 (1st cycle). Data
More importantly, its redox potential is observed at about of olivine NaFePO4 derived from Ref. [79]; data of LiFePO4 derived
3.0 V vs. Na/Na+ as compared to 4.0 V vs. Li/Li+ for LiMn2O4. from Ref. [42] b) Comparison of charge and discharge profile for
amorphous and olivine NaFePO4. Data of amorphous NaFePO4
4.1.5. LiFePO4 and NaFePO4 derived from Ref. [80] (2nd cycle).

The upper cut off potential of layered oxides is limited due


to the risk of irreversible phase transitions during cycling. On vine NaFePO4 exhibited a specific capacity of about
the other hand, the structure of polyanionic compounds is 110 mAh g@1 with a discharge voltage of about 2.8 V as seen
more stable during cycling in high operating voltages. For in Figure 10. The appearance of the additional phase also
instance, in 1997, Goodenough and his colleagues reported strongly affects the kinetic properties of the material during
the olivine-type cathode material LiFePO4, which can exhibit cycling.[82] Recently, Kim et al. reported that maricite
a reversible capacity of about 160 mAh g@1 with a working NaFePO4, despite previous expectation, can be surprisingly
voltage of about 3.5 V vs. Li/Li+.[33] In case of sodium, two electrochemically active when prepared as nanosized par-
different phases of NaFePO4, olivine and maricite, have been ticles. A reversible capacity of about 142 mAh g@1 with 95 %
considered. The maricite phase is found to be electrochemi- capacity retention after 200 cycles has been reported.[80] The
cally inactive owing to the absence of a Na diffusion path. The electrochemical activity was ascribed to a rapid amorphiza-
electrochemically active olivine NaFePO4 can be obtained by tion during initial charging. The voltage profiles for crystalline
delithiation of LiFePO4 followed by subsequent sodiation. and amorphous NaFePO4 are compared in Figure 10 b.
Figure 10 compares the voltage profiles of olivine LiFePO4
and NaFePO4, respectively. It can be clearly seen that
NaFePO4 possesses a lower redox potential as compared to 5. Materials for the Negative Electrode
LiFePO4. The most interesting difference between LiFePO4
and NaFePO4 is, however, that the latter shows a step during Before discussing different types of negative electrodes, it
charging. The plateau at 3.0 V is related to the formation of is important to remember that electrolytes are chronically
Na2/3FePO4[78] which is, in contrast to the case of lithium, instable at low potentials vs. A/A+ and decompose to form
a thermodynamically stable intermediate phase.[81] The oli- surface films. To stop electrolyte decomposition, the formed

Angew. Chem. Int. Ed. 2018, 57, 102 – 120 T 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.angewandte.org 113
Angewandte

15213773, 2018, 1, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.201703772 by Chonnam National University, Wiley Online Library on [27/12/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews Chemie

surface film must be electronically insulating but permeable


for the cations. The film should be also sufficiently thin and
flexible to survive the volume changes of the electrode during
cycling. Achieving such a stabilizing solid electrolyte inter-
phase[83] (SEI) is therefore crucial for obtaining a reversible
electrode reaction. The mechanisms of SEI formation are
very complex and are also accompanied by the formation of
soluble and gaseous compounds. The composition of the SEI
is mixed organic/inorganic and strongly depends on the used
electrolyte solvents and the conductive salt used. In LIBs, the
SEI fortunately exhibits very good properties and special
additives such as vinylene carbonate are used to tune its exact
composition. It should be noted that side reactions related to
electrolyte decomposition can be very subtle and might be
overlooked in many laboratory half-cell experiments, where
electrolyte and lithium/sodium metal are often used in excess.
It is quite predictable that SEI formation will be different
in case of sodium. The solubility of the SEI compounds in the
electrolyte will be different and the general additives known
from LIB technology might not be as effective in SIBs, for
example. Comparative studies on several electrodes so far
show that in case of sodium, the SEI is generally more
inorganic and slightly thicker as compared to the case of
lithium.[84] Overall, the knowledge on the SEI in SIBs is still
comparably small and further studies are needed to identify
the relevant factors for obtaining an optimized SEI that is
stable for a long time and over many cycles.

5.1. Graphite and Other Carbons


Figure 11. a) Intercalation of sodium and lithium into graphite (1st
Among all host structures, graphite can be considered as cycle).[94] In conventional electrolytes, only lithium intercalates, forming
quite outstanding. This is because graphite is redox-ampho- the GIC LiC6. Sodium can be intercalated by using strongly solvating
solvents. In this case, the solvated ion is intercalated (see inset).[95]
teric and the bond between the individual graphenes is weak.
EC = ethylene carbonate, DMC = dimethyl carbonate. b) Voltage pro-
Graphite can therefore intercalate not only cations, but also files for lithium/sodium insertion into hard carbon (2nd cycle), data
anions (for example, PF6@ or TFSI@)[85] of various sizes redrawn from Ref. [96].
forming a large variety of graphite intercalation compounds
(GICs).[86] With lithium, graphite forms a series of GICs at
potentials below 0.25 V vs. Li/Li+. The final composition is sodium ions into graphite using diglyme as electrolyte solvent.
LiC6, corresponding to a capacity of 372 mAh g@1.[19a] Owing Despite the large size of the solvated ion, the electrode
to its excellent overall properties, graphite is nowadays used reaction is highly reversible and exhibits excellent kinetics.
almost exclusively in commercialized LIBs. An intriguing Moreover, the redox potential can be tuned by varying the
finding is that sodium does not form any sodium-rich GICs. type of ether solvent.[89] As solvated ions are very large, one
This is indeed puzzling because GICs for for the even larger might also expect delamination of the graphite lattice.
alkali metals potassium and rubidium are well-known. The Increased van der Waals interactions between the graphene
underlying reason is still not completely clear. The increased layers and the co-intercalated solvent molecules have been
stretching of the C@C bonds of graphite during sodium recently suggested to stabilize the structure.[90] Overall, the
intercalation has been suggested by DFT calculations, for strategy of solvent co-intercalation provides new opportuni-
example.[87] ties and could also render the intercalation of other ions, for
The voltage profiles for lithium and sodium intercalation example, Mg2+ or Al3+ that otherwise do not intercalate into
into graphite are compared in Figure 11 a. Lithium intercala- graphite. On the downside, the so far obtained capacities of
tion is characterized by a staged process with the final slightly above 100 mAh g@1 are comparably low and the
formation of LiC6. The capacity is virtually zero in case of limited oxidative stability of ethers hampers their use in cells
sodium when conventional electrolyte solvents (mixture of with high voltage cathode materials.
different carbonates) are used. It was therefore commonly An alternative to graphite are disordered carbons, which
accepted that graphite cannot be used in SIBs. A strategy have been used in the early commercialized LIBs and
around this problem is to co-intercalate the solvation shell of fortunately can also store sodium when using conventional
the sodium ion as shown by Jache and Adelhelm.[88] The inset carbonate-based electrolyte solvents. The microstructure and
in Figure 11 shows the intercalation behavior of solvated composition of disordered carbons is highly dependent on the

114 www.angewandte.org T 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2018, 57, 102 – 120
Angewandte

15213773, 2018, 1, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.201703772 by Chonnam National University, Wiley Online Library on [27/12/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews Chemie

synthesis parameters and the type of precursors used. As Table 1: Overview on intermetallic lithium/sodium compounds for Zn and
a result, the shape of the voltage profiles can therefore a selection of Group 13–15 main group elements (room temperature
considerably vary. The storage behavior of lithium and phases according to the phase diagrams).[103] Li22Si5 appears in the phase
diagram but electrochemical lithiation leads to Li15Si4 only.[104]
sodium in hard carbons (hard carbons = disordered carbons
that do not transform into graphite upon heat treatment) was Li Na
compared by Stevens and Dahn.[91] The overall behavior is Zn 13 14 15 Zn 13 14 15
similar in both cases and the voltage profile consists of one LiZn AlLi Li22Si5 Li3Sb NaZn13 Ga4Na NaSi Na3Sb
sloping region and one plateau region. The former was related Li2Zn3 Al2Li3 Li15Si4 Li2Sb Ga39Na22 Ge4Na NaSb
to ion insertion between the disordered graphene layers, the (Ga13Na7)
LiZn2 AlLi2@x Li21Si8 In8Na5 GeNa
latter to ion storage by adsorption in micropores. Theoretical
Li2Zn5 Al4Li9 Li2Si InNa GeNa3
studies support this finding,[29a] although also modified con- LiZn4 Ga14Li3 GeLi3 InNa3 Na15Sn4
cepts are being suggested.[92] Overall, the storage mechanisms Ga2Li Ge5Li22 Na5Tl Na3Sn
of sodium in disordered carbons as well the impact of doping GaLi Li22Sn5 Na2Tl Na9Sn4
and nanostructuring still need to be clarified in more detail. Ga4Li6 Li7Sn2 NaTl NaSn
Figure 11 b shows a comparison for lithium and sodium Ga2Li3 Li13Sn5 NaTl2 NaSn2
insertion into hard carbon. For optimized hard carbons, GaLi2 Li5Sn2 NaSn3
InLi Li7Sn3 NaSn4
reversible capacities of around 300 mAh g@1 are obtained.[93]
In4Li5 LiSn NaSn6
Although this value is promising, it must be realized that the In2Li3 Li2Sn5 PbNa
plateau region is very close to the metal plating potential, InLi2 Li4Pb Pb4Na9
which increases the risk for undesired dendrite formation in In3Li13 Li10(8)Pb3 Pb2Na5
full cells. Other specific challenges related to sodium storage Li4Tl Li3Pb Pb4Na15
in disordered carbons are sluggish kinetics and an often too Li3Tl Li5(7)Pb2
high irreversible capacity in the first cycle. Nanosizing has Li5Tl2 LiPb
Li2Tl
become a popular approach to improve the kinetics, however,
LiTl
at the cost of an even higher initial irreversible capacity (see
for example Ref. [97]). Overall, the properties of disorder
carbons are promising but much remains to be understood
and the performance needs to be optimized further. Specif- the use of alloys in rechargeable cells. Large volume
ically, the complex interplay between carbon microstructure, expansion (often several hundred percent), insufficient SEI
redox behavior and SEI formation (electrode and electrolyte formation and a large initial irreversible capacity are the main
formulation) needs further enlightment. A recent overview drawbacks. A number of strategies such as nanosizing and
on the use of hard carbons for SIBs can be found in Ref. [98]. optimization of binder and electrolyte composition exist that
mitigate these issues, however with still too many compro-
mises. A critical discussion on the necessary conditions for
5.2. Metal Alloys increasing the energy density of lithium-ion batteries by alloys
can be found in Ref. [102]. In commercial cells, silicon is
Another promising route for storing lithium and sodium is nowadays so far only added in small quantities to graphite
alloy formation. Lithium and sodium form a number of binary electrodes.
alloys with metals of the Groups 13, 14, and 15. Full lithiation In the case of sodium-ion batteries, alloys are more and
of silicon (Li15Si4) or tin (Li15Sn4) corresponds to capacities of more under the spotlight because graphite does not show the
qth = 3578 mAh g@1 and 847 mAh g@1, which is significantly expected storage capacity (see Section 5.1). DFT calculations
higher than conventionally used electrode materials. Owing by Chevrier and Ceder showed that the average sodiation
to the large difference in electronegativity values, interme- potentials of alloys are around 0.15 V lower compared to the
tallic compounds of this kind are Zintl phases,[99] that is, they equivalent lithium reaction.[17] In case of silicon, this means
are generally brittle, often possess a stoichiometrically very that the sodiation potential for the formation of the NaSi
narrow homogeneity range (phase width) and have high intermetallic phase (qth = 954 mAh g@1) is very close to the
melting points. Table 1 shows an overview of the different metal plating potential. It might be therefore difficult to form
phases existing at room temperature. It is interesting to note this phase electrochemically. The use of amorphous rather
that, although silicon is very intensively studied for batteries, than crystalline silicon might provide advantages as suggested
the phase diagram is still a matter of debate.[100] Nevertheless, by theory.[105] Although several efforts have been under-
the number of intermetallic compounds is obviously plenty taken,[106] unequivocal evidence for the reversible storage of
and it is therefore no surprise that the lithiation and sodiation larger amounts of sodium in silicon electrodes is still missing.
processes are complex. Moreover, amorphization can occur, The sodiation and lithiation behavior of antimony (qth =
as it is known for the lithiation of silicon, for example.[101] The 660 mAh g@1) has been studied by Darwiche et al.[107] Distinct
use of alloys as negative electrodes in LIBs has been differences between both reactions where observed. In case of
comprehensively summarized by Obrovac and Chevrier.[102] lithium, the reaction proceeded in line with the phase diagram
Alloy formation with lithium takes place at potentials below over several intermediates. In case of sodium, however,
1 V vs. Li/Li+ so they are in principle well-suited as negative mostly amorphous phases as well as an unexpected high-
electrodes. A number of well-known challenges so far prevent pressure phase were observed. Interestingly, the overall

Angew. Chem. Int. Ed. 2018, 57, 102 – 120 T 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.angewandte.org 115
Angewandte

15213773, 2018, 1, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.201703772 by Chonnam National University, Wiley Online Library on [27/12/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews Chemie

performance for sodium was better than for lithium, and related to nucleation processes, so the phase formation of
capacities close to 600 mAh g@1 were achieved for more than NaSn5 is kinetically hindered.[112]
150 cycles. Specific limitations of Sb are toxicity issues as well The use binary alloys such as SbSn or even more complex
as its low abundance so research on tin and its alloys has alloys[113] are also of interest. This approach allows for a slight
become quite popular.[108] tuning in redox potentials but more importantly is often found
From charge/discharge curves, it can be seen that the to improve the cycle life because the microstructure is thought
electrochemical lithiation and sodiation of tin at room to be better preserved during cycling as compared to the pure
temperature does not proceed as one would expect from the metals. Kim et al. studied the sodiation behavior of SnSb
Li-Sn and Na-Sn phase diagrams.[108a, 109] This means that not electrodes and found reversible capacity of about
all phases shown in Table 1 necessarily form in their 500 mAh g@1 for over 100 cycles.[114] SnSb is sodiated step by
thermodynamically stable crystalline states, instead also step, where the initial reaction forms Na3Sb and Sn. After-
metastable crystalline as well as amorphous phases appear. wards Sn is further sodiated to form Na15Sn4 at lower redox
The voltage profiles for lithiation and sodiation of tin are potential.
compared in Figure 12. In both cases, a rather complex For any metal or alloy, it is also clear that sodiation will
behavior is observed. Replacing lithium by sodium leads to lead to much larger volume expansions as compared to when
lithiation takes place. Mitigating electrode degradation is
therefore expected to be more challenging in case of sodium.
Nanotomography studies on tin, however, showed that the
stability can nevertheless be better for sodium compared to
lithium.[115]
At last, trends within the periodic table can be found for
the comparison of lithium and sodium alloys. Considering the
Group 13 metals of the periodic table (Al, Ga, In, Tl), it can
be seen that lithium forms more intermetallic compounds
than sodium (Table 1). For the AB alloys of lithium, that is,
LiAl, LiGa, LiIn, LiTl, the phase width increases with
increasing atomic numbers (decreasing electronegativity).
The stoichiometric variation is around 3 % for LiAl and LiGa
and around 11 % for LiIn and LiTl, indicating that the
bonding for the latter becomes less heteropolar. Conse-
quently, also the melting point decreases with increasing
Figure 12. Voltage profiles (1st cycle) for the Na-Sn (blue) and the Li- atomic numbers. In the case of sodium, only In and Tl form
Sn system (green). Data collected from Refs. [112, 116]. AB compounds and the phase width is much smaller. The
phase width of intermetallic compounds has direct conse-
quences on the voltage profile. Phases with narrow phase
a shift in redox potentials of around 200–300 mV, making tin width lead to steps in the voltage profiles whereas a broad
a very attractive material for the negative electrode. The phase width leads to sloping potentials. The numerous
phase behavior during sodiation and desodiation of tin has intermetallic compounds for the promising Group 14 ele-
been studied by Obrovac et al.[108a] by XRD. According to ments Si and Sn show only very small homogeneity range for
their results, sodiation of tin occurs as follows, with “a” both lithium and sodium.
indicating the compound being amorphous and “*” indicating The trend that lithium forms more intermetallic com-
an unexpected crystalline phase (approximated stoichiome- pounds than sodium is interestingly opposite to what is
tries): generally found for many positive electrode materials and the
Group 16 elements oxygen, sulfur, and selenium. This might
Region 1: Na + Sn!NaSn3* well be related to the atomic radii of the alloying metals which
Region 2: Na + NaSn3*!a-NaSn are closer to lithium (rLi = 145 pm) than to sodium (rNa =
Region 3: 5 Na + 4(a-NaSn)!Na9Sn4* 180 pm). A large phase width of around 7 % is only found
Region 4: 6 Na + Na9Sn4*!Na15Sn4 for the NaPb alloy because the atomic radii are identical
(rPb = rNa = 180 pm). The large difference in atomic radii
The different phases are similar (but not identical) to what has might also explain why no Na-Al alloys exist (rAl = 125 pm).
been predicted by DFT calculations.[17] Slightly different Whether or not alloys form with lithium or sodium has
experimental results were also reported by Wang et al.,[110] important consequences for the cell design. As mentioned
indicating that the precise reaction is complex and maybe also above, aluminum can be used as current collector for the
depend on experimental conditions. A very comprehensive negative electrode in SIBs because sodium and aluminum do
study on the sodiation of tin was very recently presented by not form an alloy. In case of LIBs, more expensive copper has
Stratford et al. revealing the structure of several more to be used instead. Indium is certainly no option for
intermediate phases as well as partial solid solution behavior. commercial cells owing to its scarcity but it is worth
Eventually, even sodiation beyond the Na15Sn4 phase can mentioning that it is frequently used on the research level in
occur.[111] The peak close to the end of the desodiation is likely the rising field of solid-state batteries. It is popular because it

116 www.angewandte.org T 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2018, 57, 102 – 120
Angewandte

15213773, 2018, 1, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.201703772 by Chonnam National University, Wiley Online Library on [27/12/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews Chemie

is very soft, it readily alloys with lithium and sodium at room tinuous side reactions with the electrolyte have been found to
temperature upon contacting and the formed alloys are stable be a major obstacle,[118] so better stabilization of the electrode/
against a variety of solid electrolytes that are instable in electrolyte interface is required. Finally, it is interesting to
contact with lithium and sodium metal. note that the redox chemistry of titanates is very flexible, the
lithium titanates can store sodium ions and vice versa. Also
protonated titanates can be used.[119] TiO2, too, shows some
5.3. Li2Ti3O7 and Na2Ti3O7 peculiar behavior when comparing lithium and sodium
storage.[120]
After iron, titanium is the second most abundant tran-
sition-metal, making it an interesting element in view of
resources. The family of titanium compounds is very diverse, 6. Summary and Outlook
and plenty of compounds are being studied for use in LIBs
and more recently SIBs. Oxidic titanium compounds are Within the last 5 years, research on sodium-ion batteries
among the few being capable of reversibly intercalating has become very dynamic, and fast progress in materials
lithium and sodium at low redox potentials without under- development and performance has been achieved. There is
going the type of conversion reactions discussed in Section 3. almost no doubt that commercialization of SIBs is technically
Spinel-type Li[Li1/3Ti5/3]O4 (or often written Li4Ti5O12) has possible. This, however, requires a market niche where SIBs
been commercialized in LIBs for niche applications, for show specific advantages over LIBs or other established types
example. A particularly interesting compound for comparing of batteries. Cost-efficient batteries based on abundant
the influence of ion size on the redox behavior is A2Ti3O7 elements might be such a niche that might become relevant
(A = Li, Na), which has been studied by Rousse et al.[16] The in the future in case resource supply and supply chains for
voltage profiles are compared in Figure 13 and show a defined LIBs will be challenged too much; of course this requires an
plateau, indicative for a two-phase reaction. The redox integrated approach. Scientifically, it is intriguing to compare
potential for sodium insertion into Na2Ti3O7 (0.3 V vs. Na/ how the size of the ions affects the electrode reaction for
Na+) is more than one volt lower than compared to lithium a given host structure. In this Review, we largely focused on
insertion into Li2Ti3O7 (1.6 V vs. Li/Li+), meaning that the the differences in phase behavior during lithiation/sodiation
driving force for ion intercalation is much smaller in case of of a variety of such host structures. In many positive electrode
sodium. The difference of more than 1 V is exceptionally high materials, replacement of lithium by sodium leads to a more
compared to the other materials discussed. Moreover, the complex behavior as additional intermediate phases form
authors found that due to the different ion size, sodium during cell cycling. Layered oxides and AFePO4 (A = Li, Na)
insertion into the host structure leads to expansion, whereas have been discussed, for example. For the negative electrode,
contraction occurs in case of lithium. In situ studies during the choice so far is limited, as graphite only stores sodium
sodiation of Na2Ti3O7 confirmed that, similar to the lithium under special circumstances and silicon seems to be largely
analogue, Ti3+/4+ is the active redox couple.[117] The compound inactive. Other alloys (Sn, SnSb), titanium oxides and a variety
is able to store two sodium ions to form Na4Ti3O7 during of carbon materials are therefore intensively studied though
discharge, corresponding to a theoretical capacity of a much better understanding of the SEI formation is required.
178 mAh g@1. Poor cycle life and low initial coulombic Finally, the huge family of conversion reactions is an
efficiency, however, are the major drawbacks to date. Con- attractive playground to test whether the use of sodium
instead of lithium helps to overcome previous challenges.
Overall, numerous und unexpected surprises have been found
when replacing lithium by sodium in electrochemical cells and
much remains to be discovered. This also includes the use of
sodium in solid-state, metal/air- and metal/sulfur batteries.
The race is on to find such surprises with the hope to not only
deepen fundamental understanding but also to identify
electrode reactions for batteries with improved properties
thereby creating options for future energy storage devices.
Certainly, research on LIBs will also profit from this.

Acknowledgements

The authors thank the DFG for funding within the project
“Thermodynamics and kinetics of conversion reactions in
sodium-based battery systems” and the State of Thuringia for
support within the ProExzellenz program. L.Y. thanks the
China Scholarship Council funding. P.A. thanks M. Adelhelm
Figure 13. Voltage profiles for Li2Ti3O7 and Na2Ti3O7 (2nd cycle). Data and J. Janek for continuous and fruitful discussions.
collected from Ref. [16].

Angew. Chem. Int. Ed. 2018, 57, 102 – 120 T 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.angewandte.org 117
Angewandte

15213773, 2018, 1, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.201703772 by Chonnam National University, Wiley Online Library on [27/12/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews Chemie

Conflict of interest [19] a) M. Winter, J. O. Besenhard, M. E. Spahr, P. Novak, Adv.


Mater. 1998, 10, 725 – 763; b) M. Winter, J. O. Besenhard,
Electrochim. Acta 1999, 45, 31 – 50.
The authors declare no conflict of interest.
[20] Y. Idota, T. Kubota, A. Matsufuji, Y. Maekawa, T. Miyasaka,
Science 1997, 276, 1395 – 1397.
How to cite: Angew. Chem. Int. Ed. 2018, 57, 102 – 120
[21] P. Poizot, S. Laruelle, S. Grugeon, L. Dupont, J. M. Tarascon,
Angew. Chem. 2018, 130, 106 – 126
Nature 2000, 407, 496 – 499.
[22] a) J. Cabana, L. Monconduit, D. Larcher, M. R. Palacin, Adv.
Mater. 2010, 22, E170 – E192; b) R. Malini, U. Uma, T. Sheela,
[1] A. Thielmann, A. Sauer, M. Wietschel, Fraunhofer-Institute M. Ganesan, N. G. Renganathan, Ionics 2009, 15, 301 – 307;
ISI, Karlsruhe, 2015. c) A. Kraytsberg, Y. Ein-Eli, J. Solid State Electrochem. 2017,
[2] a) B. Nykvist, M. Nilsson, Nat. Clim. Change 2015, 5, 329 – 332; 21, 1907; d) M. Keppeler, M. Srinivasan, ChemElectroChem
b) C. Pillot, Avicenne Energy on ENSAM Meeting Paris 2016. 2017, DOI: 10.1002/celc.201700747.
[3] G. E. Blomgren, J. Electrochem. Soc. 2017, 164, A5019 – A5025. [23] F. Klein, B. Jache, A. Bhide, P. Adelhelm, Phys. Chem. Chem.
[4] C. Wadia, P. Albertus, V. Srinivasan, J. Power Sources 2011, 196, Phys. 2013, 15, 15876 – 15887.
1593 – 1598.
[24] a) A. D8bart, L. Dupont, R. Patrice, J. M. Tarascon, Solid State
[5] a) K. M. Abraham, Solid State Ionics 1982, 7, 199 – 212;
Sci. 2006, 8, 640 – 651; b) B. Jache, B. Mogwitz, F. Klein, P.
b) G. H. Newman, L. P. Klemann, J. Electrochem. Soc. 1980,
Adelhelm, J. Power Sources 2014, 247, 703 – 711.
127, 2097 – 2099; c) L. W. Shacklette, T. R. Jow, L. Townsend, J.
[25] F. Wang, R. Robert, N. A. Chernova, N. Pereira, F. Omenya, F.
Electrochem. Soc. 1988, 135, 2669 – 2674.
Badway, X. Hua, M. Ruotolo, R. G. Zhang, L. J. Wu, V. Volkov,
[6] A. Yoshino, Angew. Chem. Int. Ed. 2012, 51, 5798 – 5800;
D. Su, B. Key, M. S. Whittingharn, C. P. Grey, G. G. Amatucci,
Angew. Chem. 2012, 124, 5898 – 5900.
Y. M. Zhu, J. Graetz, J. Am. Chem. Soc. 2011, 133, 18828 –
[7] J. Peters, D. Buchholz, S. Passerini, M. Weil, Energy Environ.
18836.
Sci. 2016, 9, 1744 – 1751.
[26] M. Valvo, F. Lindgren, U. Lafont, F. Bjçrefors, K. Edstrçm, J.
[8] a) Y. Li, Y. Lu, C. Zhao, Y.-S. Hu, M.-M. Titirici, H. Li, X.
Power Sources 2014, 245, 967 – 978.
Huang, L. Chen, Energy Storage Mater. 2017, 7, 130 – 151; b) S.-
[27] F. Klein, R. Pinedo, B. B. Berkes, J. Janek, P. Adelhelm, J. Phys.
W. Kim, D.-H. Seo, X. Ma, G. Ceder, K. Kang, Adv. Energy
Chem. C 2017, 121, 8679 – 8691.
Mater. 2012, 2, 710 – 721; c) V. Palomares, P. Serras, I. Villa-
[28] A. Eguia-Barrio, E. Martinez-Castillo, F. Klein, R. Pinedo, L.
luenga, K. B. Hueso, J. Carretero-Gonzalez, T. Rojo, Energy
Lezema, J. Janek, P. Adelhelm, T. Rojo, J. Power Sources 2017,
Environ. Sci. 2012, 5, 5884 – 5901; d) H. Pan, Y.-S. Hu, L. Chen,
367, 130 – 137.
Energy Environ. Sci. 2013, 6, 2338 – 2360; e) M. D. Slater, D.
[29] a) P.-C. Tsai, S.-C. Chung, S.-K. Lin, A. Yamada, J. Mater.
Kim, E. Lee, C. S. Johnson, Adv. Funct. Mater. 2013, 23, 947 –
Chem. A 2015, 3, 9763 – 9768; b) T. B. Kim, J. W. Choi, H. S.
958; f) N. Yabuuchi, K. Kubota, M. Dahbi, S. Komaba, Chem.
Ryu, G. B. Cho, K. W. Kim, J. H. Ahn, K. K. Cho, H. J. Ahn, J.
Rev. 2014, 114, 11636 – 11682; g) D. Kundu, E. Talaie, V.
Power Sources 2007, 174, 1275 – 1278.
Duffort, L. F. Nazar, Angew. Chem. Int. Ed. 2015, 54, 3431 –
3448; Angew. Chem. 2015, 127, 3495 – 3513; h) S. Y. Hong, Y. [30] X. D. Xiang, K. Zhang, J. Chen, Adv. Mater. 2015, 27, 5343 –
Kim, Y. Park, A. Choi, N.-S. Choi, K. T. Lee, Energy Environ. 5364.
Sci. 2013, 6, 2067 – 2081; i) S. Guo, J. Yi, Y. Sun, H. Zhou, [31] X. Li, Y. Wang, D. Wu, L. Liu, S. H. Bo, G. Ceder, Chem. Mater.
Energy Environ. Sci. 2016, 9, 2978 – 3006; j) W. Luo, F. Shen, C. 2016, 28, 6575 – 6583.
Bommier, H. Zhu, X. Ji, L. Hu, Acc. Chem. Res. 2016, 49, 231 – [32] a) M. H. Han, E. Gonzalo, G. Singh, T. Rojo, Energy Environ.
240; k) L. Peng, Y. Zhu, D. Chen, R. S. Ruoff, G. Yu, Adv. Sci. 2015, 8, 81 – 102; b) H. Kim, H. Kim, Z. Ding, M. H. Lee, K.
Energy Mater. 2016, 6, 1 – 21. Lim, G. Yoon, K. Kang, Adv. Energy Mater. 2016, 6, 1 – 38.
[9] D. R. Lide, CRC Handbook of Chemistry and Physics, 84th ed., [33] A. K. Padhi, K. S. Nanjundaswamy, J. B. Goodenough, J.
CRC, Boca Raton, 2004. Electrochem. Soc. 1997, 144, 1188 – 1194.
[10] M. Okoshi, Y. Yamada, A. Yamada, H. Nakai, J. Electrochem. [34] P. Barpanda, S. Nishimura, A. Yamada, Adv. Energy Mater.
Soc. 2013, 160, A2160 – A2165. 2012, 2, 841 – 859.
[11] S. Komaba, C. Takei, T. Nakayama, A. Ogata, N. Yabuuchi, [35] D. Aurbach, K. Gamolsky, B. Markovsky, G. Salitra, Y. Gofer,
Electrochem. Commun. 2010, 12, 355 – 358. U. Heider, R. Oesten, M. Schmidt, J. Electrochem. Soc. 2000,
[12] a) P. Adelhelm, P. Hartmann, C. L. Bender, M. Busche, C. 147, 1322 – 1331.
Eufinger, J. Janek, Beilstein J. Nanotechnol. 2015, 6, 1016 – [36] K. Edstrçm, T. Gustafsson, J. O. Thomas, Electrochim. Acta
1055; b) L. Medenbach, P. Adelhelm, Top. Curr. Chem. 2017, 2004, 50, 397 – 403.
375, 81. [37] a) S. Dalavi, M. Xu, B. Knight, B. L. Lucht, Electrochem. Solid
[13] a) W. A. Hart, O. F. Beumel Jr, in The Chemistry of Lithium, State Lett. 2012, 15, A28 – A31; b) M. Xu, N. Tsiouvaras, A.
Sodium, Potassium, Rubidium, Cesium and Francium, Perga- Garsuch, H. A. Gasteiger, B. L. Lucht, J. Phys. Chem. C 2014,
mon, Oxford, 1973, pp. 331 – 367; b) A. F. Holleman, E. 118, 7363 – 7368.
Wiberg, N. Wiberg, Lehrbuch der anorganischen Chemie, de [38] D. Chen, G. Q. Shao, B. Li, G. G. Zhao, J. Li, J. H. Liu, Z. S.
Gruyter, 2007. Gao, H. F. Zhang, Electrochim. Acta 2014, 147, 663 – 668.
[14] C. H. Hamann, W. Vielstich, Elektrochemie, Wiley-VCH, [39] M. Reynaud, M. Ati, B. C. Melot, M. T. Sougrati, G. Rousse,
Weinheim, 2005. J. N. Chotard, J. M. Tarascon, Electrochem. Commun. 2012, 21,
[15] S. P. Ong, V. L. Chevrier, G. Hautier, A. Jain, C. Moore, S. Kim, 77 – 80.
X. H. Ma, G. Ceder, Energy Environ. Sci. 2011, 4, 3680 – 3688. [40] R. Tripathi, T. N. Ramesh, B. L. Ellis, L. F. Nazar, Angew.
[16] G. Rousse, M. E. Arroyo-de Domablo, P. Senguttuvan, A. Chem. Int. Ed. 2010, 49, 8738 – 8742; Angew. Chem. 2010, 122,
Ponrouch, J.-M. Tarascon, M. R. Palacin, Chem. Mater. 2013, 8920 – 8924.
25, 4946 – 4956. [41] T. Muraliganth, K. R. Stroukoff, A. Manthiram, Chem. Mater.
[17] V. L. Chevrier, G. Ceder, J. Electrochem. Soc. 2011, 158, 2010, 22, 5754 – 5761.
A1011 – A1014. [42] G. Ali, J. H. Lee, D. Susanto, S. W. Choi, B. W. Cho, K. W. Nam,
[18] M. Nose, H. Nakayama, K. Nobuhara, H. Yamaguchi, S. K. Y. Chung, ACS Appl. Mater. Interfaces 2016, 8, 15422 –
Nakanishi, H. Iba, J. Power Sources 2013, 234, 175 – 179. 15429.

118 www.angewandte.org T 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2018, 57, 102 – 120
Angewandte

15213773, 2018, 1, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.201703772 by Chonnam National University, Wiley Online Library on [27/12/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews Chemie

[43] A. Langrock, Y. H. Xu, Y. H. Liu, S. Ehrman, A. Manivannan, Appl. Mater. Interfaces 2015, 7, 3773 – 3781; c) Y. Lee, J. Lee,
C. S. Wang, J. Power Sources 2013, 223, 62 – 67. K. Y. Lee, J. Mun, J. K. Lee, W. Choi, J. Power Sources 2016,
[44] P. Singh, K. Shiva, H. Celio, J. B. Goodenough, Energy Environ. 315, 284 – 293; d) C. Chen, T. F. Geng, C. Y. Du, P. J. Zuo, X. Q.
Sci. 2015, 8, 3000 – 3005. Cheng, Y. L. Ma, G. P. Yin, J. Power Sources 2016, 331, 91 – 99;
[45] S. D. Li, J. H. Guo, Z. Ye, X. Zhao, S. Q. Wu, J. X. Mi, C. Z. e) R. B. Yu, Y. B. Lin, Z. G. Huang, Electrochim. Acta 2015,
Wang, Z. L. Gong, M. J. McDonald, Z. Z. Zhu, K. M. Ho, Y. 173, 515 – 522; f) S. W. Sun, Y. F. Yin, N. Wan, Q. Wu, X. P.
Yang, ACS Appl. Mater. Interfaces 2016, 8, 17233 – 17238. Zhang, D. Pan, Y. Bai, X. Lu, ChemSusChem 2015, 8, 2544 –
[46] C. Y. Chen, K. Matsumoto, T. Nohira, R. Hagiwara, Y. Orikasa, 2550.
Y. Uchimoto, J. Power Sources 2014, 246, 783 – 787. [64] a) J. Xu, D. H. Lee, R. J. Clement, X. Q. Yu, M. Leskes, A. J.
[47] D. M. Dai, B. Li, H. W. Tang, K. Chang, K. Jiang, Z. R. Chang, Pell, G. Pintacuda, X. Q. Yang, C. P. Grey, Y. S. Meng, Chem.
X. Z. Yuan, J. Power Sources 2016, 307, 665 – 672. Mater. 2014, 26, 1260 – 1269; b) J. Xu, H. D. Liu, Y. S. Meng,
[48] C. F. Liu, R. Masse, X. H. Nan, G. Z. Cao, Energy Storage Electrochem. Commun. 2015, 60, 13 – 16.
Mater. 2016, 4, 15 – 58. [65] D. Kim, S. H. Kang, M. Slater, S. Rood, J. T. Vaughey, N. Karan,
[49] R. Berthelot, D. Carlier, C. Delmas, Nat. Mater. 2011, 10, 74 – M. Balasubramanian, C. S. Johnson, Adv. Energy Mater. 2011,
80. 1, 333 – 336.
[50] a) H. Yoshida, N. Yabuuchi, S. Komaba, Electrochem. [66] E. de la Llave, E. Talaie, E. Levi, P. K. Nayak, M. Dixit, P. T.
Commun. 2013, 34, 60 – 63; b) X. Wang, M. Tamaru, M. Rao, P. Hartmann, H. F. Chesneau, D. T. Major, M. Greenstein,
Okubo, A. Yamada, J. Phys. Chem. C 2013, 117, 15545 – 15551. D. Aurbach, L. F. Nazar, Chem. Mater. 2016, 28, 9064 – 9076.
[51] N. Yabuuchi, H. Yoshida, S. Komaba, Electrochemistry 2012, [67] E. de la Llave, P. K. Nayak, E. Levi, T. R. Penki, S. Bublil, P.
80, 716 – 719. Hartmann, F.-F. Chesneau, M. Greenstein, L. F. Nazar, D.
[52] R. Kanno, T. Shirane, Y. Kawamoto, Y. Takeda, M. Takano, M. Aurbach, J. Mater. Chem. A 2017, 5, 5858 – 5864.
Ohashi, Y. Yamaguchi, J. Electrochem. Soc. 1996, 143, 2435 – [68] a) P. K. Nayak, J. Grinblat, M. Levi, Y. Wu, B. Powell, D.
2442. Aurbach, J. Electroanal. Chem. 2014, 733, 6 – 19; b) S. Kim,
[53] Y. S. Lee, S. Sato, Y. K. Sun, K. Kobayakawa, Y. Sato, J. Power X. H. Ma, S. P. Ong, G. Ceder, Phys. Chem. Chem. Phys. 2012,
Sources 2003, 119, 285 – 289. 14, 15571 – 15578.
[54] X. F. Wang, G. D. Liu, T. Iwao, M. Okubo, A. Yamada, J. Phys. [69] Y. C. Lei, X. Li, L. Liu, G. Ceder, Chem. Mater. 2014, 26, 5288 –
Chem. C 2014, 118, 2970 – 2976. 5296.
[55] M. Sathiya, K. Hemalatha, K. Ramesha, J. M. Tarascon, A. S. [70] S. Kumakura, Y. Tahara, K. Kubota, K. Chihara, S. Komaba,
Prakash, Chem. Mater. 2012, 24, 1846 – 1853. Angew. Chem. Int. Ed. 2016, 55, 12760 – 12763; Angew. Chem.
[56] X. F. Luo, X. Y. Wang, L. Liao, X. M. Wang, S. Gamboa, P. J. 2016, 128, 12952 – 12955.
Sebastian, J. Power Sources 2006, 161, 601 – 605. [71] J.-K. Park, Principle and applications of lithium secondary
[57] Y. K. Sun, D. J. Lee, Y. J. Lee, Z. H. Chen, S. T. Myung, ACS batteries, Wiley-VCH, Weinheim, 2012.
Appl. Mater. Interfaces 2013, 5, 11434 – 11440. [72] P.-F. Wang, Y. You, Y.-X. Yin, Y.-G. Guo, J. Mater. Chem. A
[58] E. de la Llave, V. Borgel, K. J. Park, J. Y. Hwang, Y. K. Sun, P. 2016, 4, 17660 – 17664.
Hartmann, F. F. Chesneau, D. Aurbach, ACS Appl. Mater.
[73] D. H. Lee, J. Xu, Y. S. Meng, Phys. Chem. Chem. Phys. 2013, 15,
Interfaces 2016, 8, 1867 – 1875.
3304 – 3312.
[59] a) D. Y. W. Yu, K. Yanagida, Y. Kato, H. Nakamura, J. Electro-
[74] E. Talaie, V. Duffort, H. L. Smith, B. Fultz, L. F. Nazar, Energy
chem. Soc. 2009, 156, A417 – A424; b) J. Lim, J. Moon, J. Gim,
Environ. Sci. 2015, 8, 2512 – 2523.
S. Kim, K. Kim, J. Song, J. Kang, W. B. Im, J. Kim, J. Mater.
[75] M. M. Thackeray, W. I. F. David, P. G. Bruce, J. B. Good-
Chem. 2012, 22, 11772 – 11777; c) K. Kubota, T. Kaneko, M.
enough, Mater. Res. Bull. 1983, 18, 461 – 472.
Hirayama, M. Yonemura, Y. Imanari, K. Nakane, R. Kanno, J.
[76] N. Yabuuchi, M. Yano, S. Kuze, S. Komaba, Electrochim. Acta
Power Sources 2012, 216, 249 – 255; d) S. F. Amalraj, D. Sharon,
2012, 82, 296 – 301.
M. Talianker, C. M. Julien, L. Burlaka, R. Lavi, E. Zhecheva, B.
[77] X. Z. Liu, X. Wang, A. Iyo, H. J. Yu, D. Li, H. S. Zhou, J. Mater.
Markovsky, E. Zinigrad, D. Kovacheva, R. Stoyanova, D.
Chem. A 2014, 2, 14822 – 14826.
Aurbach, Electrochim. Acta 2013, 97, 259 – 270.
[78] a) P. Moreau, D. Guyomard, J. Gaubicher, F. Boucher, Chem.
[60] a) M. M. Thackeray, S. H. Kang, C. S. Johnson, J. T. Vaughey,
R. Benedek, S. A. Hackney, J. Mater. Chem. 2007, 17, 3112 – Mater. 2010, 22, 4126 – 4128; b) J. Lu, S. C. Chung, S.-i.
3125; b) P. K. Nayak, J. Grinblat, M. Levi, D. Aurbach, Nishimura, A. Yamada, Chem. Mater. 2013, 25, 4557 – 4565.
Electrochim. Acta 2014, 137, 546 – 556; c) J. Zeng, Y. H. Cui, [79] A. Kuwahara, S. Suzuki, M. Miyayama, J. Electroceram. 2010,
D. Y. Qu, Q. Zhang, J. W. Wu, X. M. Zhu, Z. H. Li, X. H. 24, 69 – 75.
Zhang, ACS Appl. Mater. Interfaces 2016, 8, 26082 – 26090. [80] J. Kim, D. H. Seo, H. Kim, I. Park, J. K. Yoo, S. K. Jung, Y. U.
[61] H. J. Yu, H. S. Zhou, J. Phys. Chem. Lett. 2013, 4, 1268 – 1280. Park, W. A. Goddard, K. Kang, Energy Environ. Sci. 2015, 8,
[62] a) M. N. Ates, Q. Y. Jia, A. Shah, A. Busnaina, S. Mukerjee, 540 – 545.
K. M. Abraham, J. Electrochem. Soc. 2014, 161, A290 – A301; [81] F. Boucher, J. Gaubicher, M. Cuisinier, D. Guyomard, P.
b) Q. Li, G. S. Li, C. C. Fu, D. Luo, J. M. Fan, L. P. Li, ACS Moreau, J. Am. Chem. Soc. 2014, 136, 9144 – 9157.
Appl. Mater. Interfaces 2014, 6, 10330 – 10341; c) D. Wang, Y. [82] C. Heubner, S. Heiden, M. Schneider, A. Michaelis, Electro-
Huang, Z. Q. Huo, L. Chen, Electrochim. Acta 2013, 107, 461 – chim. Acta 2017, 233, 78 – 84.
466; d) P. K. Nayak, J. Grinblat, E. Levi, M. Levi, B. Markov- [83] E. Peled, J. Electrochem. Soc. 1979, 126, 2047 – 2051.
sky, D. Aurbach, Phys. Chem. Chem. Phys. 2017, 19, 6142 – [84] a) S. Komaba, W. Murata, T. Ishikawa, N. Yabuuchi, T. Ozeki,
6152; e) P. K. Nayak, J. Grinblat, M. Levi, E. Levi, S. Kim, J. W. T. Nakayama, A. Ogata, K. Gotoh, K. Fujiwara, Adv. Funct.
Choi, D. Aurbach, Adv. Energy Mater. 2016, 6, 1502398; f) C. C. Mater. 2011, 21, 3859 – 3867; b) A. Ponrouch, R. Dedryvere, D.
Wang, Y. C. Lin, P. H. Chou, RSC Adv. 2015, 5, 68919 – 68928; Monti, A. E. Demet, J. M. A. Mba, L. Croguennec, C. Mas-
g) Q. Q. Qiao, L. Qin, G. R. Li, Y. L. Wang, X. P. Gao, J. Mater. quelier, P. Johansson, M. R. Palacin, Energy Environ. Sci. 2013,
Chem. A 2015, 3, 17627 – 17634. 6, 2361 – 2369; c) B. Philippe, M. Valvo, F. Lindgren, H.
[63] a) M. Xu, Z. Y. Chen, L. J. Li, H. L. Zhu, Q. F. Zhao, L. Xu, Rensmo, K. Edstrçm, Chem. Mater. 2014, 26, 5028 – 5041;
N. F. Peng, L. Gong, J. Power Sources 2015, 281, 444 – 454; b) F. d) M. A. Munoz-M#rquez, M. Zarrabeitia, E. Castillo-Marti-
Wu, X. X. Zhang, T. L. Zhao, L. Li, M. Xie, R. J. Chen, ACS nez, A. Eguia-Barrio, T. Rojo, M. Casas-Cabanas, ACS Appl.

Angew. Chem. Int. Ed. 2018, 57, 102 – 120 T 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.angewandte.org 119
Angewandte

15213773, 2018, 1, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.201703772 by Chonnam National University, Wiley Online Library on [27/12/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews Chemie

Mater. Interfaces 2015, 7, 7801 – 7808; e) R. Mogensen, D. [106] a) C.-H. Lim, T.-Y. Huang, P.-S. Shao, J.-H. Chien, Y.-T. Weng,
Brandell, R. Younesi, ACS Energy Lett. 2016, 1, 1173 – 1178. H.-F. Huang, B. J. Hwang, N.-L. Wu, Electrochim. Acta 2016,
[85] a) W. M-rkle, N. Tran, D. Goers, M. E. Spahr, P. Nov#k, Carbon 211, 265 – 272; b) Y. Xu, E. Swaans, S. Basak, H. W. Zandber-
2009, 47, 2727 – 2732; b) G. Schmuelling, T. Placke, R. gen, D. M. Borsa, F. M. Mulder, Adv. Energy Mater. 2016, 6,
Kloepsch, O. Fromm, H.-W. Meyer, S. Passerini, M. Winter, J. 1501436.
Power Sources 2013, 239, 563 – 571. [107] A. Darwiche, C. Marino, M. T. Sougrati, B. Fraisse, L. Stievano,
[86] a) M. S. Dresselhaus, G. Dresselhaus, Adv. Phys. 1981, 30, 139 – L. Monconduit, J. Am. Chem. Soc. 2012, 134, 20805 – 20811.
326; b) Graphite Intercalation Compounds I, Springer, Amster- [108] a) L. D. Ellis, T. D. Hatchard, M. N. Obrovac, J. Electrochem.
dam, 1990. Soc. 2012, 159, A1801 – A1805; b) S. Komaba, Y. Matsuura, T.
[87] a) K. Nobuhara, H. Nakayama, M. Nose, S. Nakanishi, H. Iba, Ishikawa, N. Yabuuchi, W. Murata, S. Kuze, Electrochem.
J. Power Sources 2013, 243, 585 – 587; b) G. Yoon, H. Kim, I. Commun. 2012, 21, 65 – 68; c) M. K. Datta, R. Epur, P. Saha, K.
Park, K. Kang, Adv. Energy Mater. 2017, 7, 1601519. Kadakia, S. K. Park, P. N. Kuma, J. Power Sources 2013, 225,
[88] B. Jache, P. Adelhelm, Angew. Chem. Int. Ed. 2014, 53, 10169 – 316 – 322; d) Y. Xu, Y. Zhu, Y. Liu, C. Wang, Adv. Energy
10173; Angew. Chem. 2014, 126, 10333 – 10337. Mater. 2013, 3, 128 – 133; e) D. Bresser, F. Mueller, D. Buchholz,
[89] a) H. Kim, J. Hong, G. Yoon, H. Kim, K.-Y. Park, M.-S. Park, E. Paillard, S. Passerini, Electrochim. Acta 2014, 128, 163 – 171.
W.-S. Yoon, K. Kang, Energy Environ. Sci. 2015, 8, 2963 – 2969; [109] I. D. Courtney, J. R. Dahn, J. Electrochem. Soc. 1997, 144, 6,
b) B. Jache, J. O. Binder, T. Abe, P. Adelhelm, Phys. Chem. 2045 – 2052.
Chem. Phys. 2016, 18, 14299 – 14316; c) H. Kim, J. Hong, Y.-U. [110] J. W. Wang, X. H. Liu, S. X. Mao, J. Y. Huang, Nano Lett. 2012,
Park, J. Kim, I. Hwang, K. Kang, Adv. Funct. Mater. 2015, 25, 12, 5897 – 5902.
534 – 541. [111] J. M. Stratford, M. Mayo, P. K. Allan, O. Pecher, O. J. Borkie-
[90] S. C. Jung, Y.-J. Kang, Y.-K. Han, Nano Energy 2017, 34, 456 – wicz, K. M. Wiaderek, K. W. Chapman, C. J. Pickard, A. J.
462. Morris, C. P. Grey, J. Am. Chem. Soc. 2017, 139, 7273 – 7286.
[91] D. A. Stevens, J. R. Dahn, J. Electrochem. Soc. 2001, 148, [112] B. Zhang, G. Rousse, D. Foix, R. Dugas, D. A. Corte, J. M.
A803 – A811. Tarascon, Adv. Mater. 2016, 28, 9824 – 9830.
[92] C. Bommier, T. W. Surta, M. Dolgos, X. L. Ji, Nano Lett. 2015, [113] B. Farbod, K. Cui, W. P. Kalisvaart, M. Kupsta, B. Zahiri, A.
15, 5888 – 5892. Kohandehghan, E. M. Lotfabad, Z. Li, E. J. Luber, D. Mitlin,
[93] J. Zhao, L. Zhao, K. Chihara, S. Okada, J.-i. Yamaki, S. ACS Nano 2014, 8, 4415 – 4429.
Matsumoto, S. Kuze, K. Nakane, J. Power Sources 2013, 244, [114] Y. Kim, K. H. Ha, S. M. Oh, K. T. Lee, Chemistry 2014, 20,
752 – 757. 11980 – 11992.
[94] P. Adelhelm, Nachr. Chem. 2014, 62, 1163 – 1168. [115] J. Wang, C. Eng, Y.-C. K. Chen-Wiegart, J. Wang, Nat.
[95] See Ref. [88]. Commun. 2015, 6, 7496.
[96] M. Dahbi, N. Yabuuchi, K. Kubota, K. Tokiwa, S. Komaba, [116] J. Li, F. Yang, J. Ye, Y.-T. Cheng, J. Power Sources 2011, 196,
Phys. Chem. Chem. Phys. 2014, 16, 15007 – 15028. 1474 – 1477.
[97] a) S. Wenzel, T. Hara, J. Janek, P. Adelhelm, Energy Environ. [117] J. Xu, C. Ma, M. Balasubramanian, Y. S. Meng, Chem.
Sci. 2011, 4, 3342 – 3345; b) K. Tang, L. Fu, R. J. White, L. Yu, Commun. 2014, 50, 12564 – 12567.
M.-M. Titirici, M. Antonietti, J. Maier, Adv. Energy Mater. [118] J. Nava-Avendano, A. Morales-Garcia, A. Ponrouch, G.
2012, 2, 873 – 877. Rousse, C. Frontera, P. Senguttuvan, J. M. Tarascon, M. E. A.-
[98] a) E. Irisarri, A. Ponrouch, M. R. Palacin, J. Electrochem. Soc. D. Dompablo, M. R. Palacin, J. Mater. Chem. A 2015, 3, 22280 –
2015, 162, A2476 – A2482; b) C. Bommier, Y. Ji, Isr. J. Chem. 22286.
2015, 55, 5. [119] a) K. Chiba, N. Kijima, Y. Takahashi, Y. Idemoto, J. Akimoto,
[99] E. W. Zintl, Z. Elektrochem. 1935, 41, 876 – 879. Solid State Ionics 2008, 178, 1725 – 1730; b) A. Egu&a-Barrio, E.
[100] M. Zeilinger, I. M. Kurylyshyn, U. H-ussermann, T. F. F-ssler, Castillo-Mart&nez, M. Zarrabeitia, M. A. Munoz-M#rquez, M.
Chem. Mater. 2013, 25, 4623 – 4632. Casas-Cabanas, T. Rojo, Phys. Chem. Chem. Phys. 2015, 17,
[101] H. H. Li, X. Huang, L. Chen, G. Zhou, Z. Zhang, D. Yu, Y. Mo, 6988 – 6994.
N. Pei, Solid State Ionics 2000, 135, 181 – 191. [120] L. Wu, D. Bresser, D. Buchholz, G. A. Giffin, C. R. Castro, A.
[102] M. N. Obrovac, V. L. Chevrier, Chem. Rev. 2014, 114, 11444 – Ochel, S. Passerini, Adv. Energy Mater. 2015, 5, 1401142.
11502. [121] a) L. Oliveira et al., J. Clean. Prod., 2015, 108, 354 – 362; b) J. F.
[103] M. E. Schlesinger, E. M. Mueller, Alloy Phase Diagrams, ASM Peters, M. Weil, Resources, 2016, 5, 46.
Handbook Volume 3, ASM International, Materials Park, OH,
1983. Manuscript received: April 12, 2017
[104] J. Li, J. R. Dahn, J. Electrochem. Soc. 2007, 154, A156 – A161. Revised manuscript received: June 9, 2017
[105] S. C. Jung, D. S. Jung, J. W. Choi, Y.-K. Han, J. Phys. Chem. Lett. Accepted manuscript online: June 19, 2017
2014, 5, 1283 – 1288. Version of record online: November 20, 2017

120 www.angewandte.org T 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2018, 57, 102 – 120

You might also like