ArticlePDF Available

Practical Cathodes for Sodium‐Ion Batteries: Who Will Take The Crown?

Wiley
Advanced Energy Materials
Authors:

Abstract and Figures

In recent decades, sodium‐ion batteries (SIBs) have received increasing attention because they offer cost and safety advantages and avoid the challenges related to limited lithium/cobalt/nickel resources and environmental pollution. Because the sodium storage performance and production cost of SIBs are dominated by the cathode performance, developing cathode materials with large‐scale production capacity is the key to achieving commercial applications of SIBs. Therefore, developing host materials with high energy density, long cycling life, low production cost, and high chemical/environmental stability is crucial for implementing advanced SIBs. Among the developed cathode materials for SIBs, O3‐type sodiated transition‐metal oxides have attracted extensive attention owing to their simple synthesis methods, high theoretical specific capacity, and sufficient Na content. However, the relatively large Na‐ion radius leads to sluggish diffusion kinetics and inevitable complex phase transitions during the deintercalation/intercalation process, resulting in poor rate capability and cycling stability. Therefore, this review comprehensively summarizes the research progress and modification strategies for O3‐type cathodes, including the component design, surface modification, and optimization of synthesis methods. This work aims to guide the development of commercial layered oxides and provide technical support for the next generation of energy‐storage systems.
This content is subject to copyright. Terms and conditions apply.
REVIEW
www.advenergymat.de
Practical Cathodes for Sodium-Ion Batteries: Who Will Take
The Crown?
Xinghui Liang, Jang-Yeon Hwang, and Yang-Kook Sun*
In recent decades, sodium-ion batteries (SIBs) have received increasing
attention because they offer cost and safety advantages and avoid the
challenges related to limited lithium/cobalt/nickel resources and
environmental pollution. Because the sodium storage performance and
production cost of SIBs are dominated by the cathode performance,
developing cathode materials with large-scale production capacity is the key
to achieving commercial applications of SIBs. Therefore, developing host
materials with high energy density, long cycling life, low production cost, and
high chemical/environmental stability is crucial for implementing advanced
SIBs. Among the developed cathode materials for SIBs, O3-type sodiated
transition-metal oxides have attracted extensive attention owing to their
simple synthesis methods, high theoretical specific capacity, and sufficient Na
content. However, the relatively large Na-ion radius leads to sluggish diffusion
kinetics and inevitable complex phase transitions during the
deintercalation/intercalation process, resulting in poor rate capability and
cycling stability. Therefore, this review comprehensively summarizes the
research progress and modification strategies for O3-type cathodes, including
the component design, surface modification, and optimization of synthesis
methods. This work aims to guide the development of commercial layered
oxides and provide technical support for the next generation of energy-storage
systems.
1. Introduction
At present, lithium-ion batteries (LIBs) successfully dominate
the market of portable electronic devices, and their applica-
tion is gradually expanding into the field of electric vehicles.[]
Nevertheless, the limited and uneven distribution of lithium
(Li), nickel (Ni), and cobalt (Co) worldwide has raised concerns
about the sustainable development of this battery technology.
Sodium-ion batteries (SIB) are considered the most appropriate
X. Liang, J.-Y. Hwang, Y.-K. Sun
Department of Energy Engineering
Hanyang University
Seoul 04763, Republic of Korea
E-mail: yksun@hanyang.ac.kr
J.-Y. Hwang, Y.-K.Sun
Department of Battery Engineering
Hanyang University
Seoul 04763, Republic of Korea
The ORCID identification number(s) for the author(s) of this article
can be found under https://doi.org/10.1002/aenm.202301975
DOI: 10.1002/aenm.202301975
alternative to LIBs because of their abun-
dant raw materials, potentially low cost,
and similar intercalation chemistry.[,]
In fact, the development of both SIBs and
LIBs originated in the s. However,
with the discovery of the Li-storage mech-
anism of graphite anodes in the s,
research on SIBs stagnated because
sodium (Na) has a large ionic radius and
cannot be eectively inserted/extracted
from graphite anodes. During this pe-
riod, high-temperature Na batteries (ZE-
BRA cells) using molten Na as the anode
and ceramic as the separator developed
rapidly, but the high operating tempera-
ture (– °C) led to issues such as
low energy eciency and corrosion. With
the development of hard-carbon (HC)
anodes and rising concerns about Li re-
sources and the supply chain of Li/Co/Ni
raw materials, SIB development has
accelerated since  (Figure 1a).
Similar to Li technology, an SIB com-
prises a cathode, anode, separator, and
electrolyte.[] Benefiting from the low
cost of cathodes and the possibility of re-
placing the expensive copper (Cu) foils
used as anode current collectors with
inexpensive aluminum (Al) foils, the total cost of SIBs is reported
to be at least % lower than that of LIBs.[] Due to the relatively
high standard electrochemical potential of Na metal (. V
versus standard hydrogen electrode (SHE) for Na; . V ver-
sus SHE for Li), Na-ion full cells can be fully discharged to V
for storage without structural degradation, which improves their
safety during operation and transportation. However, since the
energy density of SIBs is generally lower than that of LIBs, the
real unit energy cost of the two technologies is almost equal
($. per W h for SIBs and $. per W h for LIB).[] Admit-
tedly, it would be unfair to directly compare Li technology, which
has been commercialized since , with Na technology, which
is still in the prototype stage. Based on various advanced elec-
trode materials developed in recent years, the energy density
and cycle life of SIBs are expected to be further improved.[, ]
For instance, Faradion Limited pointed out that the produc-
tion cost of their second-generation cathode (Faradion’s Gen
material: mixed O-Na.Ni. Mn.Mg. Ti. Oand P-
Na.Ni.Mn. Mg.Ti.O)//HC Na-ion pouch cells will
be –% lower than that of the current state-of-the-art
LiFePO/graphite Li-ion cells.[] Therefore, considering that
Adv. Energy Mater. 2023, 2301975 © 2023 Wiley-VCH GmbH
2301975 (1 of 41)
www.advancedsciencenews.com www.advenergymat.de
Figure 1. a) Number of publications on SIBs and cathode materials for SIBs. b) Proportion of publications on various cathode materials between
2013 and 2023 (data collected using Web of Science, May 16, 2023). c) Typical crystal structures of layered oxides (O3-type), polyanionic compounds
(NASICON), and PBAs (cubic phase structure). d) Key properties of layered oxide, polyanionic compounds, and Prussian blue analogs.
large-scale energy storage systems and moderate-range electric
vehicles require long cycling life and high safety, cost-eective
SIBs have received renewed attention in recent years.[]
To accelerate the commercialization of SIBs, several startup
companies, including Faradion Limited, HiNa, CATL, Altris,
and Tiamat, have made significant progress.[] In most scale-
up and demonstration cases, prototype Na-ion pouch cells are
typically assembled using an HC anode, but the optimal cath-
ode material is yet to be identified. CATL (China), Natron En-
ergy (USA), and Novasis Technologies (USA) chose Prussian
Blue analogs as the development direction for SIBs; Faradion
Limited (UK) developed a series of doped layered oxide cath-
odes (NaaNi(xyz)MnxMgyTizO); Tiamat (France) produced
cylindrical SIBs with a NaV(PO)F(NVPF) cathode; and
HiNa (China) fabricated a  kWh battery device in Liyang
city based on a Nax[Cu,Fe,Mn]Ocathode.[,–] The practi-
cal large-format prototype SIBs shows similar electrochemi-
cal performance to the early commercial LIBs. For example,
a  A h prismatic NaCrO//HC full cell can deliver a cell
energy density of  W h kg(calculated based on all cell
Adv. Energy Mater. 2023, 2301975 © 2023 Wiley-VCH GmbH
2301975 (2 of 41)
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202301975 by Hanyang University Library, Wiley Online Library on [18/08/2023]. 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
www.advancedsciencenews.com www.advenergymat.de
components), with a capacity retention of % after  cy-
cles at  A.[] Na.Cu.Ni.Fe.Mn. Ti. O//HC[] and
NaNi. Zn.Mn. Ti. O//HC[] Na-ion full cells are able to
deliver practical energy densities of  and  W h kgin A h
cylindrical cells (-type cells), respectively. The  A h pouch
cell can be realized by coupling HC as anode and Faradion’s Gen
material as cathode, with a high cell energy density of  W
hkg
and an ultra-long service life of  cycles.[] Last year,
a A h pouch cell (anode-free Na battery) was developed, uti-
lizing Na[Cu/Ni/ Fe/ Mn/]Oas the cathode and Al foil as
the anode, exhibiting a very high cell energy density of  W h
kg.[] Although the heavy steel casing of prismatic and cylindri-
cal cells reduces the practical energy density, the sti metallic cas-
ing and safety valve enhance the safety of the battery. Considering
that SIBs are mainly used for stationary energy storage systems
rather than portable electronic devices, prismatic and cylindrical-
type SIBs with higher safety may be more suitable for commer-
cial applications. Therefore, developing advanced cathode mate-
rials and constructing practical full cells are crucial for achieving
commercial applications of SIBs. This review paper specifically
focuses on the challenges, solutions, and future development di-
rections encountered by SIBs in large-scale applications.
2. Characterization and Selection of Suitable
Cathode Materials
The ideal cathode material should possess the advantages of
high capacity, suitable operating voltage (outputting high en-
ergy density without decomposing the electrolyte), high power
density, sucient electronic/ionic conductivity, and high chemi-
cal/environmental stability. To ensure the stability and safety of
SIBs, commercial cathode materials should also have the advan-
tages of low environmental impact, easy preparation, abundant
raw materials, and high thermal stability.[] To date, researchers
have developed various types of cathode materials, including
Prussian Blue analogs (PBAs), polyanionic compounds, organic
compounds, and sodiated transition metal oxides (Figure b).[,]
Although the advantages of structural diversity, high theoretical
capacity, and low environmental impact of organic compounds
have attracted the attention of researchers, their large-scale ap-
plications are hindered by slow reaction kinetics and the dissolu-
tion of active materials.[] In addition, based on the current de-
velopment path of SIB production companies in the field of cath-
ode materials, it is reasonable to believe that the first commercial
cathode materials will probably be PBAs, polyanionic polymers,
and/or layered oxide materials.[] Therefore, the basic principles
of these three potential inorganic cathode materials are discussed
below to critically evaluate the development direction of SIB com-
mercialization.
2.1. Polyanionic Compounds
Benefiting from the stable framework structure and strong in-
ductive eects, polyanionic compounds are considered promis-
ing cathode materials for advanced SIBs. Polyanionic com-
pounds are covalently linked through (XO)n(X =sulfur
(S), phosphorus (P), silicon (Si), and other elements) groups
and generally provide three-dimensional (D) Na-ion diusion
channels (Figure c).[] Currently, common research systems
include phosphates (NaFePO,Na
V(PO),NaVOPO
), py-
rophosphates (NaMPO,Na
MPO,Na
M(PO)PO)(M=
manganese (Mn), Co, Ni, iron (Fe), vanadium (V) and oth-
ers), fluorophosphates (Na(VOx)(PO)Fx,Na
MPOF), sul-
fates (NaFe(SO),Na
+xMx(SO),NaMSO
F), and silicates
(NaMnSiO,Na
FeSiO).[] Compared with layered oxides, a D
framework structure can eectively alleviate the structural rear-
rangement and inhibit the dissolution of oxygen (O) during the
Na+insertion/extraction process, thereby enabling good cycla-
bility and thermal stability.[] However, these cathode materials
possess inherently low electrical conductivity due to their unique
structure. Taking NaV(PO)(NVP) as an example, since O
atoms are shared in the adjacent VOoctahedra and POtetrahe-
dron, electron transfer follows the V–O–P–O–V pattern instead
of the faster V–O–V pattern.[] Therefore, to improve the intrin-
sically low electronic conductivity, conductive carbon (C) coat-
ings, nanostructure designs, and elemental doping have been
proposed.
As a typical Na superionic conductor (NASICON)-type cath-
ode, NVP has attracted widespread attention from researchers
due to its high structural/thermal stabilities and ion conduc-
tivity. NVP cathodes with a rhombohedral structure can pro-
vide a theoretical capacity of  mA h gand exhibit a flat
charge/discharge plateau at . V (versus Na+/Na), correspond-
ing to the redox pairs of V+/V+.[] Although NVP has many ad-
vantages as a cathode material, its inherently low electronic con-
ductivity (<Scm
) limits its application potential. Gener-
ally, highly conductive C-layer coatings are considered the most
eective materials to improve the electrochemical performance
of NVP. Therefore, amorphous-C, graphite-like C, porous C,
S/nitrogen (N)/P/boron (B)-doped C, carbon nanotubes (CNTs),
and graphene-modified NVP have been evaluated. The use of
these materials resulted in enhanced cycling stability and rate
properties, because the introduced C-coating not only increases
the electronic conductivity but also acts as a buer layer to pre-
vent damage to the active materials.[, ] In addition, shorten-
ing the ion-transport path by changing the microstructure of the
cathode material is an eective method to improve the rate per-
formance and long-term cycling stability of NVP. Hence, a se-
ries of cathodes with nanofiber, micro-flower, and microsphere
morphologies has been reported, which exhibited ultra-high cy-
cle lives (over  cycles).[, ] Moreover, heteroatom doping
(with Al, Mn, magnesium (Mg), zinc (Zn), molybdenum (Mo),
niobium (Nb), lanthanum (La), and other elements) has been
proven to eectively enhance the electrochemical performance of
NVP, as the doping agent expands the Na-ion diusion channel,
produces additional vacancies, and reduces the ionic migration
barrier.[] Through the eorts of researchers, modified NVP ma-
terials have exhibited ultra-long cycling stability of over  
cycles and ultra-high rate capability of above  C (although
some researchers criticize this excessively high rate as meaning-
less in practice), which is even better than the performance of
most reported cathode materials for LIBs.[, ] Notably, owing to
the diculty in activating V+/V+redox pairs, NVP-based cath-
odes can only deliver a limited energy density, and the high con-
tent of expensive and toxic V limits its potential for large-scale
production.
Adv. Energy Mater. 2023, 2301975 © 2023 Wiley-VCH GmbH
2301975 (3 of 41)
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202301975 by Hanyang University Library, Wiley Online Library on [18/08/2023]. 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
www.advancedsciencenews.com www.advenergymat.de
Since the introduction of strong electronegative groups (such
as fluorine ion (F)) can increase the inductive eect, fluorinated
phosphates with relatively high operating voltage and energy
density are attractive cathode candidates for advanced SIBs.[]
Swoyer et al. first reported the NaVPOF cathode with a tetrag-
onal structure (space group I4/mmm), which exhibited a spe-
cific capacity of  mA h gand an average discharge voltage of
. V.[] The C-coated monoclinic NaVPOF(spacegroupC2/c)
prepared by a solution-based approach exhibited a specific capac-
ity of  mA h gand excellent cycling stability ( cycles).[]
An in situ X-ray diraction (XRD) study of the phase evolution
and reaction kinetics of tetragonal and monoclinic NaVPOF
showed that, unlike the monoclinic structure that undergoes a
two-phase reaction, the tetragonal NaVPOF undergoes a single-
phase reaction during the charge/discharge process, resulting
in reduced structural damage and improved electrochemical
performance.[] Although the optimized NaVPOF exhibits im-
proved Na storage performance, its actual capacity is lower than
the theoretical value. It was found that single-phase NaVPOFis
dicult to achieve by solid-phase synthesis (NaF:VPO=:),
indicating that the crystal structure of NaVPOF requires fur-
ther investigation.[] As an important member of the V-based
fluorophosphate family, Na(VOxPO)F+x ( x) has re-
ceived extensive attention. A typical NVPF consists of VOFbi-
octahedral and POtetrahedron units and exhibits a high aver-
age operating voltage (. V) and stable structure.[] Croguen-
nec et al. re-evaluated the crystal structure of NVPF and found
that the structural dierences in the reported literature may come
from fluctuations in the ratio of F and O.[] Synchrotron radia-
tion diraction studies revealed a small amount of orthorhom-
bic distortion in NVPF, implying that the appropriate crystal
space group should be Amam rather than the commonly assigned
P42/mnm. To overcome the inherently low electronic conduc-
tivity of fluorophosphates, researchers successfully encapsulated
NVPF in a C-matrix using various methods (hydrothermal, elec-
trospinning, spray-drying, sol-gel, and ball-milling methods), re-
sulting in unique morphologies with short Na-ion diusion paths
that accelerate electron transport.[] For example, an N-doped C-
coated nanocube-like NVPF cathode material was produced by a
surfactant-assisted solid-state method, which exhibited excellent
rate capability (. mA h gat  C) and long-cycle stability
(.% capacity retention after  cycles).[] Furthermore, a C-
free high-entropy cathode (NaV.(Ca,Mg,Al,Cr,Mn). (PO)F)
material was developed, which enhanced the electronic conduc-
tivity, reduced ionic diusion barriers, and increased the average
discharge capacity (. V) compared to the original NVPF.[] As
a result, this high-entropy cathode exhibited higher energy den-
sity (from . to . W h kg) and demonstrated ultra-long
cycling stability (maintaining .% of the initial capacity after
 cycles). It should be noted that harmful F-containing gases
generated during the synthesis process may limit the application
of fluorophosphate-based cathodes.
To reduce the content of V and increase the working volt-
age, metal-ion substitution was proposed. For example, a series
of NASICON-structured cathode materials, NaxMV(PO)(M =
Fe, Mn, Ni), were developed.[] Among them, NaMnV(PO)
(NMVP) with high operating voltage (. and . V plateaus)
emerged as a promising cathode candidate and displayed higher
energy density and lower cost than NVP. However, the rate per-
formance and long-term cycling stability of NMVP are limited
by its low electronic conductivity and the Jahn–Taller (JT) dis-
tortion of Mn+, which can be improved through C-coating,
atomic doping, or nanostructuring. For example, a graphene
aerogel-coated NMVP cathode material with a high energy den-
sity of  W h kgand an ultra-long cycling life of over
 cycles was achieved.[] Furthermore, an N-doped C-coated
bayberry-like NMVP cathode was fabricated by ball-milling.[]
The conductive C-coating and hierarchical nanostructure pro-
moted ion/electron migration, and the optimized cathode exhib-
ited a high reversible capacity (. mA h g) and a long cy-
cling life (.% capacity retention after  cycles). In addi-
tion, V-free phosphates have attracted attention owing to their
low cost. These include Mn-based cathode materials with high
potentials (. V for Mn+/Mn+;. V for Mn+/Mn+), such
as NaMnTi(PO)(NMTP), NaMnZr(PO),Na
MnAl(PO),
and NaMnCr(PO).[] Based on the charge compensation of
the Mn+/+/+redox reaction, these cathode materials typically
exhibit limited reversible capacity. For example, NMTP exhib-
ited a low discharge capacity of  mA h gwithin a volt-
age window of .–. V and poor cycling stability (the dis-
charge capacity dropped to  mA h gafter  cycles).[] Al-
though a C-coating can improve the reversible capacity (e.g.,
 mA h g) and energy density (e.g.,  W h kg) of NMTP,
its service life is lower than that of most V-based polyanionic com-
pounds (.% capacity retention after  cycles at C).[] Al-
though reducing the discharge cuto voltage of NMTP can ac-
tivate the titanium (Ti+/Ti+) redox couple at . V to provide
a high theoretical capacity of  mA h g, its abnormal initial
Coulombic eciency (ICE) and low-voltage plateau lead to low
practicability.[]
Given the higher thermal stability of pyrophosphate (PO)an-
ions than phosphate anions, structurally stable pyrophosphates
are another strong contender for high-performance cathode ma-
terials. Yamada’s group first reported a novel cathode material for
SIBs, namely NaFePOwith a triclinic structure, which exhib-
ited an average operating voltage of V and a reversible capac-
ity of  mA h g.[] Dierential scanning calorimetry results
showed that the desodiated NaFePOmaintained structural in-
tegrity up to  °C, demonstrating its good thermal stability.[]
A novel Mn-based sodium pyrophosphate (NaMnPO) cathode
was shown to provide a reversible capacity of  mA h gand an
average discharge voltage of . V.[] Despite pyrophosphate
cathode materials exhibiting high structural/thermal stabilities,
the low specific capacity caused by heavy anionic groups may hin-
der their commercial application. Therefore, mixed phosphates
consisting of phosphate (PO)and pyrophosphate (PO)an-
ions have received increasing attention. For example, potentially
low-cost NaFe(PO)POand NaMn(PO)POhave been
presented, which exhibited high reversible capacities ( and
 mA h g, respectively) and high average discharge voltages
(. and . V, respectively).[ ,] Although both cathode mate-
rials exhibit high energy density, their low electronic conductiv-
ity, JT distortion, and dissolution of active materials result in un-
satisfactory cycling stability, which requires further optimization.
In conclusion, polyanionic compounds with high structural sta-
bility are considered as potential commercial cathode materials,
but further optimization of the component design and synthe-
sis methods is needed to reduce production costs (reducing the
Adv. Energy Mater. 2023, 2301975 © 2023 Wiley-VCH GmbH
2301975 (4 of 41)
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202301975 by Hanyang University Library, Wiley Online Library on [18/08/2023]. 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
www.advancedsciencenews.com www.advenergymat.de
content of expensive V) and increase the volumetric energy den-
sity (C-matrix and nanostructure).
2.2. Prussian Blue Analogs
Prussian blue analogs (PBAs) have been widely studied be-
cause of their facile preparation methods and potentially low
cost.[] PBAs are generally represented by the formula of
AxM[M’(CN)]yy·nHO(x; y <), where repre-
sents the vacancies occupied by coordinating HO, A represents
alkali metal ions (Li, Na, potassium (K)), and M/M’ is Fe, Mn,
Ni, Co, Zn, or chromium (Cr).[] Among them, hexacyanofer-
rates (Fe(CN), HCFs) have become the most attractive subgroup
due to their abundant resources and high redox potential. In the
case of Na-based materials, the content of Na ions in the unit
cell aects the color (blue, green, or white) and crystal struc-
ture (cubic, rhombohedral, or monoclinic phase) of the result-
ing cathode material, leading to dierent electrochemical perfor-
mances. Fully sodiated Prussian white is considered the most
suitable cathode material because this Na-rich phase structure
is conducive to the commercial application of Na full cells.[]
Compared with sodiated transition metal oxides and phosphates,
PBAs with larger interstitial A sites show higher ionic diusion
coecients (to cms), allowing reversible electro-
chemical reactions of large Na ions.[] According to the redox re-
action, PBAs can be divided into single-electron transfer types (M
=Zn, Ni, Cu, Cr, and M’ =Fe, Co, Mn) and double-electron trans-
fer types (M and M’ =Fe, Co, Mn). PBAs with single-electron re-
actions exhibit negligible lattice parameter changes (almost zero
strain) during repeated charging/discharging processes, provid-
ing long-term cycling stability.[] However, the low theoretical
specific capacity of  mA h glimits their application poten-
tial. PBAs with two-electron reactions exhibited a high theoret-
ical capacity of  mA h g; for example, Na.Fe[Fe(CN)]
with a rhombohedral structure provided a reversible capacity of
 mA h gand an average operating voltage of . V.[] Due
to lattice distortion, the monoclinic phase undergoes a complex
phase transformation (from monoclinic to cubic, then tetragonal
phase), resulting in poor cycling stability and low Coulombic e-
ciency. Because the cubic structure with high symmetry is main-
tained during the electrochemical process, the development of
cubic-phase PBAs with high Na contents can eectively increase
the structural stability of the modified cathode. Accordingly, cu-
bic Na.Mn[Fe(CN)].·.HO was synthesized using a high
concentration of trisodium citrate as a chelating agent at  °C.[]
The cubic phase is typically present in PBAs with low Na content,
as interstitial water enters the crystal lattice when the Na-ion con-
tent increases, resulting in the formation of monoclinic or rhom-
bohedral structures with higher amounts of interstitial water.[]
Therefore, compared to the monoclinic MmHCF, the resulting
cubic MnHCF with slightly higher interstitial water and slightly
lower Na contents sacrificed some reversible capacity (from .
to . mA h g).[] However, its high structural reversibility
during the electrochemical process provided excellent rate prop-
erties (. mA h gat  mA g) and cycling stability (%
capacity retention after  cycles at  mA g). Optimizing the
composition of active/inactive metal ions to balance the trade-o
between capacity and structural stability is another eective strat-
egy to enhance the Na-storage performance of PBAs. The struc-
tural stability and ionic diusion of Fe/Mn-based PBAs can be
improved by introducing inactive elements. However, excessive
substitution results in capacity loss (e.g., Zn, Cu, Ni) or higher
production costs (e.g., Co).[] For example, high concentrations
of Zn substitution lead to the transition of Na.Fe[Fe(CN)]from
the rhombohedral to the cubic phase.[] Due to the improved
structural reversibility, the cycling stability of the optimized cath-
ode (Na.Zn.Fe.[Fe(CN)]) was greatly improved (.% ca-
pacity retention after  cycles at  mA g), but at the cost of
some capacity (. mA h gfor FeHCF and . mA h gfor
ZnFeHCF). Another study showed that a small amount (. per
unit) of Zn-ion substitution can maintain the monoclinic struc-
ture of NaxFe[Fe(CN)].[] Due to the reduced ionic diusion bar-
rier and enhanced activity of low-spin Fe+, the modified cathode
exhibited high reversible capacity ( mA h g) and cycling sta-
bility (% capacity retention after  cycles at A g). In ad-
dition, single/dual element-doped and high-entropy PBAs have
been shown to improve the Na-storage performance, indicating
that appropriate compositional design can improve the perfor-
mance of PBAs to some extent.[, ]
Another key factor limiting the electrochemical performance
of PBAs is the presence of vacancies and defects in the frame-
work structure.[] During PBA fabrication by the precipita-
tion method, the fast precipitation reaction inevitably introduces
M’(CN)defects in the crystal structure. These defects are sponta-
neously occupied by water molecules to form coordinated water.
The vacancies and coordinated water in the structure can cause
lattice distortion and structural damage during cycling, leading
to the accumulation of capacity loss. In addition, the dissociated
crystal water undergoes significant side reactions with the elec-
trolyte and releases large amounts of gas. This poses a huge safety
concern for pouch-type full cells.[] Therefore, it is necessary to
put more eort into developing PBAs with high crystallinity, high
Na content, fewer defects, and minimal interstitial water. The
crystal water in the unit cell of PBAs mainly comprises adsorbed,
interstitial, and coordinated water. Adsorbed and interstitial wa-
ter can be easily removed through common drying procedures.
However, the elimination of strongly bound coordinated water re-
quires higher drying temperatures, which can damage the crys-
tal structure.[] Thus, the development direction for PBAs is
to minimize the presence of defects without damaging the in-
tegrity of the crystal structure. It was previously demonstrated
that rhombohedral NaMnFe(CN)·zHO, from which intersti-
tial water can be removed through vacuum drying ( °C), ex-
hibits a high reversible capacity of  mA h gand a combined
charging/discharging plateau (about . V).[] PBAs with low in-
terstitial water were synthesized on a large scale using a sim-
ple heat-treatment method (dried in an argon (Ar) atmosphere
at  °C for h).[] In situ high-temperature synchrotron XRD
analysis revealed a new triangular structure in the resulting cath-
ode, which activated the low-spin Fe+/+redox reaction, resulting
in excellent cycling stability (.% capacity retention after 
cycles at C).
To further reduce the amounts of vacancies and inter-
stitial water in the crystal structure, it has been proposed
to use high-concentration Na reaction solutions or chelating
agent/surfactant-assisted coprecipitation methods. Increasing
the concentration of Na ions in the reaction solution allows more
Adv. Energy Mater. 2023, 2301975 © 2023 Wiley-VCH GmbH
2301975 (5 of 41)
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202301975 by Hanyang University Library, Wiley Online Library on [18/08/2023]. 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
www.advancedsciencenews.com www.advenergymat.de
Na ions to fill the interstitial sites preferentially and is an eec-
tive way to reduce the amount of coordinated water. The lattice
parameters of the PB cathode increase with increasing NaCl con-
centration, indicating that more Na-ions enter the crystal unit.[]
Thermogravimetric analysis results showed that by reducing the
precipitation reaction rate, the content of interstitial water de-
creased from . to .% (although it remained at a high level),
indicating that this could be an eective way to synthesize per-
fect crystals.[] Due to the rapid nucleation and crystal growth
processes, it is dicult to synthesize PBAs with controllable size
and few defects. Therefore, researchers proposed introducing
chelating agents that bond with metal ions to reduce the co-
ordination reaction rate.[] For example, citric acid and citrate
salts, commonly used as chelating agents, not only reduce va-
cancies and interstitial water but also serve as reducing agents
to prevent the oxidation of Fe+, thereby increasing the Na con-
tent in the crystal lattice. Hence, Mn-PBA cathodes with high
Na content and few vacancies were synthesized using saturated
trisodium citrate, and regular spherical microparticles (. μm)
were obtained after aging at  °C for  h.[] The modified
cathode with a smooth surface and high tap density exhibited a
high reversible capacity of  mA h gand good cycling sta-
bility (the capacity retention increased from  to .% after
 cycles at  mA g). Because high synthesis temperatures
may increase the risk of defect formation, an ice-assisted synthe-
sis method was proposed to reduce the crystal growth rate.[]
The results of inductively coupled plasma and element analy-
sis showed that defects in the final product reduced from 
to %. In addition, other surfactants (such as polyvinylpyrroli-
done) and chelating agents (including ascorbic acid, acetic acid,
and sodium ethylene diamine tetraacetate) were evaluated to im-
prove the crystallinity of PBAs.[] Wang et al. proposed the prin-
ciple of balanced coordination to design PBAs with less intersti-
tial water and fewer transition-metal ion vacancies.[] A modi-
fied cathode (Na.Fe[Fe(CN)].·. HO) with .% intersti-
tial water was synthesized using sodium carboxymethylcellulose
as a suitable chelating agent. It exhibited high reversible capac-
ity ( mA h g) and ultra-stable cycling performance (capac-
ity retention over % after  cycles). Moreover, the addi-
tion of immiscible organic solvents to water to create a salt-in-
water-in-oil environment was successfully applied to the synthe-
sis of high-crystallinity PBA materials.[, ] Na–x FeFe(CN)was
synthesized with a low crystal water content (.%) by a low-
temperature solvothermal method with a mixture of water and
ethylene glycol.[] By adjusting the solvent ratio and reaction
time, the size of monodispersed cubic particles was controlled
( μm), and the optimized sample exhibited high reversible ca-
pacity ( mA h g) and excellent cycling stability (.% ca-
pacity retention after  cycles at A g).
The particle size is another important factor for commercial
cathode materials, as it aects the capacity, volumetric energy
density, and productivity. Due to the rapid precipitation reaction
rate, PBA materials typically exhibit a nanocubic morphology
with a large specific surface area that facilitates the activation
of charge-transfer reactions, thereby increasing the reversible
capacity.[,] However, developing nanoscale PBA materials is
impractical because of their low crystallinity and insucient tap
density. From a practical perspective, micro-sized PBAs with low
surface area and high tap density can inhibit electrolyte decom-
position and side reactions, thereby extending the service life of
the full cell. Therefore, synthesizing PBAs with fewer defects
by reducing the coordination reaction rate can increase particle
size and enhance their applicability.[,,] High-entropy single-
crystal PBAs ( μm) were prepared by a modified coprecipitation
method to reduce the number of grain boundaries to suppress
surface side reactions and increase the chemical/thermal stabil-
ity by the use of micro-sized particles.[] The optimized cathode
exhibited excellent electrochemical performance, with a high ca-
pacity retention rate of .% after  cycles at  mA g.
Microspherical PB was assembled from cubic primary particles
by spray drying and then coated with reduced graphene oxide to
improve its ionic conductivity.[] The low content of coordinated
water and improved electronic conductivity resulted in an opti-
mized cathode with a high reversible capacity of . mA h g
(activation of low-spin Fe) and excellent cycling stability (.%
capacity retention after  cycles). In conclusion, PBAs that can
be synthesized at low temperatures have good cycling stability
and cost advantages, but issues related to their limited thermal
stability, electronic conductivity, and tap density, and excessive
defects need to be resolved.
2.3. Layered Oxides
Sodiated transition-metal oxides (NaxTMO;x; where TM is
a transition metal such as Ni, Co, Mn, Fe, Cu, V, or Cr) are con-
sidered promising cathode materials for SIBs because of their
high specific capacity, easy synthesis, and good electrochemical
performance.[] According to the coordination environment of
Na ions (trigonal prismatic site or octahedral site) and the num-
ber of TMOlayers in a repeating stacking unit, Na layered ox-
ides are mainly classified into P (ABBA oxide ion stacking) and
O (ABCABC oxide ion stacking) types.[] The P-phase cath-
ode materials generally exhibit excellent cycling stability over a
narrow voltage window (e.g., .–. V versus Na+/Na) but suf-
fer from rapid capacity fading over a wider voltage range (e.g.,
.–. V).[] The instability of the P-type cathode over a wide
potential window mainly originates from the severe phase transi-
tion (P to O structure) at . V and the dissolution of TM ions
at low voltages. More importantly, due to the low Na content of
P-type cathode materials, they generally exhibit low first-charge
capacity and abnormal ICE, greatly hindering the large-scale ap-
plication of Na full cells.[] Therefore, some P-type cathode ma-
terials with Na vacancies exhibit a promising high energy den-
sity of  W h kgin Na half-cell configurations.[] Neverthe-
less, only % of this is retained in the Na-ion full cell due to
the initial Na deficiency.[] Although a pre-sodiation treatment
can compensate for the irreversible capacity loss during the first
charge, this step increases the manufacturing cost of the full cell,
reducing the potential technical advantages of SIBs.[] There-
fore, considering the success of layered oxide cathodes in LIBs,
O-type cathodes that can be paired with an HC anode to fab-
ricate practical Na-ion full cells with high specific capacity and
long cycling life are considered the most promising candidate
for commercialization.[] The challenges and corresponding so-
lutions for O-type cathode materials are discussed in the next
section.
Adv. Energy Mater. 2023, 2301975 © 2023 Wiley-VCH GmbH
2301975 (6 of 41)
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202301975 by Hanyang University Library, Wiley Online Library on [18/08/2023]. 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
www.advancedsciencenews.com www.advenergymat.de
Figure 2. a) Average operating voltage, specific capacity, and energy density of three representative cathodes. b) Cycle life of three kinds of cathodes.
2.4. Commercialization Potential
Owing to their cost and safety advantages, SIBs have attracted
attention from both academia and industry. However, there is a
gap between these two fields needs to be bridged through further
research eorts. Therefore, the commercialization prospects of
the candidate cathode materials mentioned above are discussed
considering their electrochemical performance, cost, safety, and
environmental impact. Figure 2 shows the capacity, gravimetric
energy density, and cycling stability of dierent cathode materi-
als. The gravimetric energy density of layered oxides is higher
than that of PBAs and polyanionic compounds (Figure a). More
importantly, the volumetric energy density of layered oxides is
much higher than that of polyanionic compounds and PBAs
(low atomic packing density).[] The energy density of the cath-
ode material has a significant impact on the unit energy cost of
Na-ion full cells because more active materials and electrolytes
are required to compensate for a low energy density. Conversely,
polyanionic compounds exhibit outstanding cycling stability ow-
ing to their stable framework structure (Figure b). Fortunately,
it is feasible to optimize the electrochemical performance of the
cathode materials by specific strategies to meet practical require-
ments. For example, the structural/air stability of layered oxides
can be improved by elemental substitution, surface modification,
and structural optimization strategies; the reversible capacity and
cycling stability of PBAs can be enhanced by optimizing the
Adv. Energy Mater. 2023, 2301975 © 2023 Wiley-VCH GmbH
2301975 (7 of 41)
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202301975 by Hanyang University Library, Wiley Online Library on [18/08/2023]. 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
www.advancedsciencenews.com www.advenergymat.de
composition, reducing defects, and increasing the particle size;
the inherently low electronic conductivity of polyanionic com-
pounds can be overcome by surface C-coating and nanostructur-
ing.
Although the Na-ion storage performance of the modified cath-
ode can be improved through component optimization, the com-
position of transition metals greatly aects the price of the final
product, so it is necessary to consider a balance between cost and
performance.[] Reducing the manufacturing cost of the cathode
is the key to the successful commercialization of secondary bat-
teries, especially for SIBs with potential low cost as the selling
point. Although the cost of Na is particularly low, its content in
cathode materials is limited (i.e., especially for polyanionic com-
pounds with large molecular weight); therefore, the manufactur-
ing cost of the cathode is more related to the raw material cost of
the cations in the framework. Introducing expensive metal ions
such as Co, Li, and rare-earth elements greatly increase the pro-
duction cost of the cathode materials, making them unpopular
in SIB systems. Therefore, it is necessary to reduce or even elim-
inate the usage of expensive metals (generally, to below %).
Due to the widespread use of inexpensive Fe/Mn ions, PBA cath-
odes exhibit the lowest production cost (when evaluated only on
raw material costs). For layered oxides, although the cost of Ni
is higher than that of Mn and Fe, the average operating voltage
and energy density of Ni-based layered oxides are relatively high,
resulting in a significant advantages in energy cost (lower than
LiFePO) for layered oxides composed of a small amount of Ni
(/) and other non-precious metals (such as Mn, Mg, Fe, Al,
Cu).[] In contrast, V-based polyanionic compounds are the most
mature technology and exhibit satisfactory cycling stability and
rate performance in SIBs. Nevertheless, the usage of expensive
and toxic V limits their potential for large-scale application. To
this end, researchers have developed low-cost (Fe/Mn-based) or
high-energy density (pyrophosphates, fluorophosphates) polyan-
ionic compounds to reduce energy costs. However, their Na stor-
age performance needs further optimization. In terms of pro-
cessing cost, unlike polyanionic compounds and layered oxides
that need to be calcined at high temperature to obtain the ideal
crystal structure, PBA cathodes that can be synthesized by copre-
cipitation without further processing once again show unparal-
leled cost advantages.[] Although the synthesis of layered oxides
through coprecipitation method increases the processing cost of
the cathode material to a certain extent, the regular morphology
increases the tap density, and the scale eect reduces the actual
production cost, making the processing cost of layered oxides and
polyanionic polymers acceptable. On the other hand, it is neces-
sary to develop newer electrode slurry combinations to simulta-
neously reduce costs and improve electrochemical performance.
In particular, solvent-free casting technology (dry method) have
been proposed to increase the energy density as well as reduce the
cost of electrode processing even though it was still early stage.[]
Considering that the energy density of SIBs is inevitably lower
than that of LIBs, optimizing the active materials and synthesis
process to further reduce the cost of SIBs is the key to its indus-
trialization.
Scaling up the production capacity is another important chal-
lenge faced by commercial cathode materials and is related to
the synthesis methods. The coprecipitation process suitable for
synthesizing PBAs and layered oxide cathodes oers significant
advantages in batch consistency and large-scale preparation po-
tential. This is because it involves the homogenization of the liq-
uid phase, which has been fully demonstrated in LIBs systems.[]
However, synthesizing PBAs still needs further optimization,
as the rapid precipitation reaction rate results in materials with
small particle sizes and numerous defects.[] Polyanionic com-
pounds are synthesized by solid-state, hydrothermal, sol-gel, elec-
trospinning, and spray-drying methods, among which the solid-
state method is widely used because of its simple operation, low
cost, and few by-products.[] However, solid-state methods are
typically used in research studies for synthesizing small amounts
of products, making it necessary to consider the consistency of
the cathode materials when these methods are up-scaled.
The environmental impact of commercial cathode materials
is another important factor. The by-products produced during
the synthesis of layered oxides mainly include ammonium salt,
water, and CO, which have little harmful impact on the en-
vironment. However, the synthesis of PBAs generates a large
amount of wastewater containing high concentrations of Na salts,
chelating agents/surfactants, and highly toxic cyanide (CN),
which could result in serious environmental problems.[] There-
fore, the ecient elimination or reuse of this wastewater is an-
other challenge for the commercialization of PBAs. Polyanionic
compounds containing toxic V are harmful to the environment,
while fluorophosphates with high energy density release toxic hy-
drogen fluoride (HF) gas during high-temperature calcination.
Therefore, polyanionic compounds seem to have a higher envi-
ronmental impact than layered oxides.
Finally, safety is crucial for the commercialization of SIBs.
Considering that the thermal runaway of SIBs is usually triggered
by material degradation, it is necessary to evaluate the thermal
stability of the cathode, especially under abusive operating con-
ditions such as overcharging, fast charging, and ultra-high tem-
perature cycling. Benefiting from their strong covalent bonds,
polyanionic compounds exhibit high thermal stability.[] Layered
oxides generate free Ounder high voltage or ultra-high temper-
ature (– °C), accompanied by heat release, thus exhibiting
moderate thermal stability.[] In the case of PBAs, in addition
to the side reactions caused by interstitial water, the cyanide re-
leased during their decomposition above  °C also reacts with
the electrolyte and releases heat, demonstrating the lowest safety
of the three types of cathode materials.[] In summary, layered
oxides are assumed to have the highest overall commercialization
potential because of their favorable electrochemical performance,
cost, scalability, environmental impact, and safety (Figure d).
3. O3-tye Layered Oxide Cathodes
As a Na reservoir, O-type cathodes are considered suitable can-
didates for advanced SIBs because of their high capacity and
potentially low production costs. However, there are still some
scientific issues and significant drawbacks that need to be ad-
dressed before they can be commercialized. Therefore, this re-
view summarizes the main challenges of O-type layered oxide
cathode materials for SIBs and outlines the current mainstream
optimization strategies (composition, structure, and interface) to
bridge the gap between academia and industry (Figure 3).
Adv. Energy Mater. 2023, 2301975 © 2023 Wiley-VCH GmbH
2301975 (8 of 41)
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202301975 by Hanyang University Library, Wiley Online Library on [18/08/2023]. 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
www.advancedsciencenews.com www.advenergymat.de
Figure 3. Schematic diagram of the main challenges and optimization strategies for O3-type layered oxide cathode materials.
3.1. Challenges of O3-Type Na-TM-O Cathodes
3.1.1. Sluggish Na-Ion Diffusion Kinetics
The Na/vacancy ordering caused by the electrostatic repulsion
of Na ions, JT distortions, and the charge ordering of TM ions
results in a multi-step voltage plateau and a reduced ionic dif-
fusion coecient.[, ] Due to the strong in-plane repulsion of
Na+–Na+, the transformation of Na/vacancy ordering is more
significant than that of Li/vacancy ordering.[] In addition, com-
pared with the P-type cathode with an open ion-diusion path,
because the direct hop from adjacent octahedral sites in the O-
type layered system requires high activation energy, Na ions pref-
erentially enter face-shared interstitial tetrahedral coordination
sites temporarily. Both theoretical simulations and experimental
measurements have revealed that in Na/Fe/ Mn/ Ocathodes,
the Na ion diusion coecient of the O-type structure is lower
than that of the P phase structure (Figure 4a), which greatly hin-
ders the rate capability.[] In addition, the changes in the stacking
sequence caused by electrochemical creep aect the Na-ion mi-
gration mechanism.[] A systematic investigation of the eect of
Co substitution on the electrochemical performance and struc-
tural evolution of the O-NaFeOsystem showed that the intro-
duction of Co ions maintains the P-type phase structure dur-
ing the wide operating voltage window, which is the key factor to
achieving the high rate capability of O-NaFe/Co/O.[] How-
ever, the detrimental phase transition at high potential causes ad-
verse Na migration behavior, which needs to be optimized by ion
doping or multiphase structural design.[, ] For example, first-
principles calculations showed that Mg-ion doping in NaCoO
causes charge disproportionation and the formation of electron
holes, thus significantly reducing the diusion barrier and im-
proving the rate capability.[] In addition, strategies such as par-
ticle nanosizing and morphological control have been proposed
to improve the diusion kinetics of Na ions. Because the char-
acteristic time for diusion (𝜏) is proportional to the square of
the diusion length (L;𝜏=L/D,whereDis the diusion coe-
cient of Na ions), adjusting the morphology to shorten Lis one of
the eective methods to enhance the rate properties of cathode
materials.[]
3.1.2. Deleterious Phase Transitions
A first-principles study showed that the O-type Na layered oxides
have no thermodynamic driving force to transform into spinel
structures because the relatively large size of Na ions leads to an
increase in the TM migration barrier.[] For instance, NaxTMO
oxides were synthesized from spinel LiMnOby electrochemi-
cal Li/Na ion exchange. It was demonstrated that the insertion
of Na ions into the spinel phase introduces lattice strain and fur-
ther transformation into a new monoclinic layered structure.[]
These findings show that the Na-ion system can compete with the
Li-ion system, providing an optimistic perspective for developing
Adv. Energy Mater. 2023, 2301975 © 2023 Wiley-VCH GmbH
2301975 (9 of 41)
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202301975 by Hanyang University Library, Wiley Online Library on [18/08/2023]. 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
www.advancedsciencenews.com www.advenergymat.de
Figure 4. a) Ion-diffusion energy barriers in the P2 and O3 phase structures. Reproduced with permission.[88] Copyright 2017, Wiley-VCH GmbH.
b) In situ XRD pattern of Na[Ni0.5Mn0.5 ]O2during the initial charging process. c) Cross-sectional scanning electron microscopy (SEM) images of
Na[Ni0.5Mn0.5 ]O2under different charging states. Reproduced with permission.[107 ] Copyright 2020, Wiley-VCH GmbH. d) Schematic diagram of the
structural degradation and crack formation of Na0.8Mg0.2 Fe0.4 Mn0.4O2due to exposure to air. Reproduced with permission.[108 ] Copyright 2021, the Royal
Society of Chemistry. e) Schematic diagram of chemically induced surface reconstruction. Reproduced with permission.[109] Copyright 2018, Wiley-VCH
GmbH.
cathode materials with a stable structure and high capacity.[]
However, the extraction of Na ions induces the sliding of the
TMOlayers at a relatively low potential to minimize electrostatic
repulsion and the Na-ion diusion barrier, which may modify the
coordination environment of Na.[] In addition, the JT eect of
high-spin TM (Mn(III), Ni(III), Fe(IV), and Cu(II)) induces local
structural distortion and lattice mismatch. Specifically, the octa-
hedral MnOcomplex contains high-spin Mn(III) with the asym-
metric occupation state of the egorbital (t2g3eg1), which implies
that the overall dorbital does not match with the octahedral Oh
symmetry, resulting in the instability of Mn(III) cations.[] To
stabilize trivalent Mn cations, the octahedral unit deforms with
the elongation of two longitudinal Mn–O bonds and the shrink-
age of the other four horizontal Mn–O bonds, which leads to
the lattice deformation from a hexagonal phase to an orthogonal
structure.[] Although first-principles calculations showed that
the O’ and P’ structures with JT distortion are dynamically
stable, such geometric deformation causes structural degrada-
tion and deficient electrochemical performance.[ ] As shown in
Figure b, extracting Na ions from a typical O-NaNi. Mn. O
cathode leads to the reduction of the screening eect between
the Oanion planes and further induces the lattice distortion
and phase transition of Ohexagonal to O’monoclinic to Phexagonal to
P’monoclinic to P’hexagonal to O’hexagonal.[ –] Such multi-step
phase transitions have been widely observed in many O-type
cathode materials.[, ] A multi-step phase transition that re-
quires additional energy consumption is usually accompanied by
large volume changes and collapse of the crystal structure, result-
ing in large polarization and poor cycling stability. A novel O-
type Na cathode (NaxLix[(Mn.Co.Ni.). .]Oy)was
synthesized from Li-rich layered oxides through Li/Na electro-
chemical exchange.[ ] The introduction of TM and O vacancies
induces the reversible migration of TM ions between the TM slab
and Na layer, thus increasing the kinetic barrier of the O-P
Adv. Energy Mater. 2023, 2301975 © 2023 Wiley-VCH GmbH
2301975 (10 of 41)
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202301975 by Hanyang University Library, Wiley Online Library on [18/08/2023]. 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
www.advancedsciencenews.com www.advenergymat.de
phase transition, which maintains the O-type structure during
the electrochemical process (.–. V). However, considering
that the electrochemical Li/Na exchange method is not suitable
for commercial application, it is necessary to further explore the
reaction mechanism of O-type Na layered oxides and propose
eective strategies to inhibit adverse phase transitions.
3.1.3. Mechanical Failure
The phase transformation that occurs during electrochemi-
cal processes is also accompanied by the anisotropic shrink-
age/expansion of the unit cell, resulting in the accumulation of
residual stress along grain boundaries.[ ] It has been shown that
electrochemically induced mechanical degradation is responsible
for the capacity degradation of intercalation-layer cathodes.[ ]
First, the Na concentration gradient formed during the electro-
chemical process leads to non-uniform lattice deformation. The
internal tension caused by the expansion of the c-parameter leads
to delamination cracks, while the compression caused by the con-
traction of the a-axis induces bending and kinking. These lattice
defects accumulate during cycling, resulting in slow reaction ki-
netics and degradation of the active materials. For instance, the
phase transition from hexagonal P’ to hexagonal O’ above . V
was shown to be the main reason for the rapid capacity decay
of O-NaNi.Mn.Ocathodes.[ ] It was proven that the phase
transformation accompanied by the anisotropic changes in the
lattice parameters led to the inhomogeneous accumulation of in-
ternal stress on the secondary particles and the formation of mi-
crocracks extending to the surface (Figure c). Although the in-
tegrity of secondary particles can be restored during the discharge
process, the repeated opening and closing of cracks cause perma-
nent damage to the cathode materials. The microcracks through
the secondary particles accelerate the penetration of the corro-
sive electrolyte, resulting in the continuous depletion of the elec-
trolyte and chemical damage (formation of electrochemical in-
active NiO-like rock-salt phase). Although the cycling stability of
Na layered oxides can be improved to some extent by limiting
the voltage window, this constrains the average operating voltage
and reversible capacity of SIBs.[ ] Fortunately, it is possible to
solve this problem by delaying/inhibiting the structural transfor-
mation by elemental substitution and relieving the mechanical
stress by optimizing the microstructure.
3.1.4. Surface Residual Species and Surface Degradation
The air sensitivity of layered oxides, especially sodiated TM ox-
ides with large interlayer spacing, imparts strict requirements
for the production, storage, and transportation of cathode ma-
terials, which poses a huge obstacle to the large-scale application
of SIBs.[ ] When the layered oxides are exposed to humid air,
for even a short time, some of the active Na migrates to the sur-
face and reacts with HO/COto generate residual alkali species,
such as NaOH, NaHCO,orNa
CO.[ ] The loss of active Na
accompanying the oxidation of TM ions results in the sliding
of the TMOlayer, resulting in irreversible structural changes
and significant capacity loss. An in-situ mass spectrometry study
showed that the exposure of the O-Na.Mg.Fe.Mn.Ocath-
ode to air triggers the formation of spinel- and rock salt-like struc-
tures, accompanied by the evolution of oxygen gas (O).[ ] The
gas pressure caused by Orelease forms internal cracks and elec-
trochemical failure (Figure d). In addition, water molecules are
easily inserted into the Na layer and form a hydrated phase after
Na extraction, which greatly hinders the ionic diusion channel
and increases the interfacial resistance.[ ] This problem is made
worse by the electronic/ionic insulating residual alkali species
that reduce the electronic conductivity of the cathode material
and trigger the dehydrofluorination reaction and cross-linking
of the polyvinylidene fluoride binder, leading to slurry gelation,
particle agglomeration, uneven coating, and corrosion of the cur-
rent collector, gas generation.[] Moreover, the irreversible loss
of the active material due to corrosion by HF (as a reaction by-
product) cannot be ignored. Therefore, the air stability of layered
oxides is crucial for the large-scale application of SIBs. A fea-
sible ethanol-washing method was proposed to remove surface
residues, thus improving the electrochemical performance of O-
NaNi. Mn.O.[ ] Although this post-treatment method signif-
icantly reduces the surface impurities on the cathode material, it
increases the number of production steps and waste-treatment
costs. Hence, it is necessary to further optimize the structure of
layered oxides to fundamentally improve their air stability.
Researchers have proposed surface engineering and elemental
doping strategies to improve the environmental stability of lay-
ered cathodes. A protective coating can be introduced to act as a
physical barrier for sensitive cathode materials to prevent direct
contact between the active material and humid air. A phosphate-
coating strategy was proposed to improve the air stability of O-
NaNi. Fe. Mn.O(capacity loss of  mA h gafter exposure
to humid air for days).[ ] In addition, reducing the interlayer
spacing of the Na-layer and enhancing the Na–O bonds to prevent
the insertion of HO/COmolecules is another eective strategy
for improving environmental stability. The introduction of appro-
priate Na vacancies (Na.Li.Ni.Fe.Mn. O)wasshown
to enhance the oxidation resistance of TM ions and lower the dif-
fusion energy barrier, thereby simultaneously improving air sta-
bility and reaction kinetics.[ ] The structural and air stability of
Na[Ni. Mn.]Owas enhanced by calcium (Ca) doping, as the
strong Ca–O bond inhibits the Na+/HO+exchange.[]
During electrochemical processes, especially at high voltage,
the cathode material and electrolyte undergo parasitic oxidation
reactions, which result in the continuous consumption of the
electrolyte and reconstruction of the surface crystal structure.[]
It is generally believed that the relatively large radius of Na
ions makes the migration of TMs in the Na system kinetically
unfavorable.[] However, it has been shown that when the lay-
ered oxide cathode is in a highly charged (desodiated) state, metal
ions in the TMOlayer can migrate into the Na layer, especially
Cr and Fe.[] It was pointed out that there was no Cr migration
before the transformation from O-NaCrOto P’-Na.CrO,
but further removal of Na (charging over . V) triggers the
phase transition from layered to rock-salt phase (P’-Na.CrO
to O-Na𝛿CrOto rock-salt CrO).[ ] The migration of Fe-ions
from the TMOslab to the Na layer was also observed at the
atomic scale.[ ] During desodiation, Fe ions are activated into
the JT-active high-spin Fe+state to provide structural deforma-
tion to accommodate Fe migration and reduce the energy bar-
rier of Fe penetrating the O-triangle, resulting in inevitable Fe
Adv. Energy Mater. 2023, 2301975 © 2023 Wiley-VCH GmbH
2301975 (11 of 41)
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202301975 by Hanyang University Library, Wiley Online Library on [18/08/2023]. 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
www.advancedsciencenews.com www.advenergymat.de
migration.[, ] Although the migration of TM ions is partially
reversible, it comes at the cost of irreversible capacity loss and
high polarization. Considering that Fe in octahedral sites prefer-
entially migrates to Na octahedrons through tetrahedron sites in
the Na layer, the absence of tetrahedral sites in the P-type struc-
ture may prevent TM migration to some extent.[, ] However,
the formation of Z-or OP-phase structures by stacking O- and
P-type structures provides the necessary Fe migration environ-
ment under deep desodiation. Therefore, inhibiting structural
deformation without reducing the concentration of active ions
is the developmental goal for advanced layered oxides.[ ] Re-
cently, it was proposed that the random extraction of Na and the
subsequent formation of the so-called Na-free layer at high po-
tential are the main reasons for TM migration.[ ] A ruthenium
(Ru)/Ti co-doping strategy was further developed to achieve a se-
lective desodiation reaction through the pre-oxidation of active
4d Ru, thus inhibiting the formation of a Na-free layer. More-
over, this irreversible TM migration results in the formation of
a random and uneven reconstructed layer composed of spinel
and/or rock-salt phases on the surface of the cathode material,
accompanied by oxygen loss.[ ] The anisotropic reconstructed
layer blocks ion diusion and forms edge dislocations with the
ordered Na-ion diusion layer, greatly aecting the Na-ion trans-
port kinetics. Furthermore, the dissolution of TM ions (for exam-
ple, the disproportionation reaction of Mn+Mn++Mn+)
and the accompanying nano-cracks are also responsible for ca-
pacity losses.[ ] The dissolution of TMs results in the collapse
of the layered structure, accompanied by the migration and depo-
sition of these metals on the anode, which act as catalysts for elec-
trolyte decomposition.[ ] Therefore, the described above surface
chemical process is the key reason for the inferior electrochemi-
cal performance of layered oxides (Figure e).
To date, crystal structural design, elemental substitution, mor-
phological control, and surface engineering have been demon-
strated as eective strategies to improve the structural stability,
reaction kinetics, and thermal/environmental stability of layered-
oxide cathode materials.[, ] Therefore, this section discusses
feasible development directions for O-type cathodes based on
published research work and the requirements of commercial
SIBs.
3.2. Potential O3-Type Layered-oxide Cathodes
Notably, despite the many similarities between Li and Na tech-
nologies, SIBs cannot be developed by simply using the Na coun-
terpart of LIB cathode materials. For example, the similar ionic
radii of Li and Fe (rLi/rFe =.) easily lead to cation mixing,
resulting in the formation of a cubic rock-salt 𝛼-LiFeOphase
rather than an electrochemically active layered structure. Surpris-
ingly, 𝛼-NaFeO(rNa /rFe =., hexagonal layered structure) ex-
hibits suitable Na storage activity with a distinct charge/discharge
plateau around . V.[ ] Similarly, LiCrOis considered an elec-
trochemical inactive cathode material because oxidized Cr+ions
irreversibly migrate to the Li layer and occupy interstitial tetra-
hedral sites, thereby blocking ionic diusion channels.[ ] How-
ever, the mismatched ionic radius and tetrahedron height result
in Cr remaining in the TMOslab within a narrow voltage win-
dow (below . V), so the NaCrOcathode provides a discharge
capacity of  mA h gand an average operating voltage of V.
In , Cu was introduced into a Ni/Mn-based cathode mate-
rial (Na.[Cu.Fe.Mn.]O), with the electrochemical activity
of the Cu+/Cu+redox couple evaluated using X-ray absorption
near-edge spectroscopy (XANES) analysis.[ ] The incorporation
of active Cu ions into the TMOlayer simultaneously improves
the structural stability, reaction kinetics, and air stability, provid-
ing more opportunities for the large-scale application of layered
oxides. It should be noted that, unlike LIBs with a high-Ni and
Co-free cathode, the optimal compositions of the cathode materi-
als for SIBs have not yet been determined.[ ] Based on the high
voltage provided by Fe and Cr, high capacity provided by Ni, high
electronic conductivity provided by Co, and good structural sta-
bility provided by Mn and Cu, researchers have developed a se-
ries of single (NaCrO), binary (Mn/Ni and Mn/Fe-based oxides),
ternary (MnNiCo, NiMnFe and FeMnCu systems), and multi-
metal oxides.[ ]
3.2.1. NaCrO2
O-NaCrOcathodes with excellent thermal and cycling sta-
bility have received widespread attention from researchers.[ ]
Ex-situ XRD results revealed that NaCrOundergoes a simple
phase transition from hexagonal O to monoclinic O and fi-
nally to monoclinic P during desodiation (.–. V), endow-
ing it with excellent structural reversibility (Figure 5a).[ ] Large-
grained NaCrOparticles were synthesized by directly calcining
NaCrO·HO at  °C.[ ] Reducing the specific surface area
to minimize side reactions and exposing the () plane to im-
prove the ionic diusion kinetics resulted in a cathode with ex-
cellent cycling stability (capacity retention of .% after  cy-
cles at C). In addition, due to the low oxidation state of Cr in the
layered structure, NaCrOneeds to be sintered in a reducing or
inert atmosphere, which provides the possibility of applying a C-
coating in situ. Fu et al. first synthesized a C-coated NaCrOcath-
ode via a simple solid-state reaction, which delivered a capacity re-
tention of % after  cycles.[ ] Hexagonal plate-like NaCrO
was synthesized by emulsion drying method, followed by C-
coating with pitch as the raw material (Figure b).[] The highly
crystalline C-layer was evenly coated on the surface of the active
material, which endowed the obtained cathode with high ionic
conductivity (Scm
) and thermal stability. Cathodes based
on NaCrOnanowires were obtained by electrospinning.[] This
unique D porous network composed of nanocrystalline sub-
units provides a fast diusion channel for Na ions while reduc-
ing the diusion resistance of ionic species, thereby improving
the electrochemical kinetics and enhancing structural stability
(Figure c). Therefore, the obtained cathode material exhibited
ultra-high rate capability (. mA h gat  C) and a wide op-
erating temperature range ( to  °C). In addition, elemen-
tal substitution strategies have been proposed to enhance the
crystal stability of layered frameworks, thereby suppressing high-
pressure phase transitions and increasing reversible capacity.[]
It was proposed that the introduction of an antimony ion, Sb+
(more than . mol), with a fixed oxidation state and high elec-
tronegativity, would lead to thermodynamically unfavorable Cr
migration.[ ] Therefore, the optimized O-Na.Cr.Sb. O
cathode undergoes a reversible O–P phase transition over a
Adv. Energy Mater. 2023, 2301975 © 2023 Wiley-VCH GmbH
2301975 (12 of 41)
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202301975 by Hanyang University Library, Wiley Online Library on [18/08/2023]. 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
www.advancedsciencenews.com www.advenergymat.de
Figure 5. a) Charge/discharge curves and phase-transition diagram of NaCrO2. b) Schematic diagram of the synthesis of C-coated plate-like NaCrO2
cathodes. Reproduced with permission.[136 ] Copyright 2015, the Royal Society of Chemistry. c) Transmission electron microscopy (TEM) images of
nanowire-like samples. Reproduced with permission.[139 ] Copyright 2019 American Chemical Society. d) Operando XRD patterns of the original (right)
and Sb-doped (left) cathode materials. Reproduced with permission.[140 ] Copyright 2022 Elsevier.
wide voltage range of .–. V without forming the irreversible
O’ phase (Figure d), resulting in a high reversible capacity
(. mA h g) and long-term cycling stability (.% capac-
ity retention after  cycles at C). Although NaCrOexhibits
excellent electrochemical performance, its toxicity and high pro-
duction cost indicate that it is not the preferred choice for com-
mercial applications.
Other single-metal oxides have significant drawbacks that hin-
der their commercialization, such as the complicated structural
evolution of NaNiO, JT distortion of NaMnO, high cost of
NaCoO, and Fe migration in NaFeO.[] To alleviate the inher-
ent drawbacks of single-metal oxides, the general research strat-
egy toward developing ideal cathode materials is to design multi-
metal oxides with synergistic eects to enhance structural stabil-
ity and alleviate phase transitions.
3.2.2. NaNi0.5Mn0.5 O2
Ni/Mn-based oxides are the most extensively studied layered
cathode materials because the Ni+/Ni+redox couple provides
high voltage and reversible capacity, while the JT-inactive Ni+
and Mn+form a robust framework structure. NaNi.Mn.O
is a classic O-type cathode material, which delivers a high re-
versible capacity of . mA h gwith .% ICE within a
potential window of .–. V.[ ] However, the specific capac-
ity of this material drops rapidly due to severe phase transi-
tions and mechanical stress accumulated during cycling. There-
fore, compositional modification has been proposed to improve
the cycling performance and rate capability of O-type cathodes.
Heterogeneous-atom substitution provides the following ben-
efits: ) increases the binding energy between TM and O to
improve the structural reversibility and reduce the ionic diu-
sion barrier;[,] ) adjusts the interlayer spacing of the Na
layer to improve ion-diusion kinetics and/or air stability;[, ]
and ) suppresses the JT eect of Mn+through charge com-
pensation (such as Mg+,Zn
+,Cu
+,andLi
+).[– ] For ex-
ample, it was proposed that Al doping can shorten the bond
length of TM–O and expand the interlayer spacing of the Na
layer by increasing the TM–O bond energy, thereby improv-
ing the cycling reversibility and ionic diusion kinetics of
the resulting cathode (NaAl.Ni. Mn.O).[  ] The modified
NaNi. Mn.Ocathodes doped with Zr,[ ] Ti,[ ] tin (Sn),[]
and Sb[ ] also demonstrated similar results. A novel O–
Na.Li. Ni. Mn.Owcathode material was synthesized by
electrochemical Na/Li-ion exchange, which provided an ultra-
high reversible capacity of  mA h gand a total energy den-
sity of  W h gwithin a voltage window of .–. V.[  ] In
addition, the introduction of non-toxic Ca+into the Na-layer has
been proposed.[, ] Due to the strong bonding between Ca and
O ions, which alleviates the sliding of the TMOslab and the
loss of Na/O, the resulting cathode (Na.Ca. [Ni.Mn. ]O)
Adv. Energy Mater. 2023, 2301975 © 2023 Wiley-VCH GmbH
2301975 (13 of 41)
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202301975 by Hanyang University Library, Wiley Online Library on [18/08/2023]. 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
www.advancedsciencenews.com www.advenergymat.de
Figure 6. Contour plots of in-situ XRD patterns of a) Sn-doped (4.0 V high-cutoff voltage) and b) Mg/Ti co-substituted (4.5 V high-cutoff voltage)
NaNi0.5Mn0.5 O2cathodes. Reproduced with permission.[159 ] Copyright 2021, American Chemical Society. Reproduced with permission.[150] Copyright
2021, the Royal Society of Chemistry. In situ XRD patterns of c) NaNi0.5Sn0.5O2(2.8–4.0 and 2.8–4.2 V in the initial and second cycles, respectively) and
d) NaNi0.5Mn0.25 Ti0.25O2(2.0–4.0 V). Reproduced with permission.[ 162 ] Copyright 2018, Wiley-VCH GmbH. Reproduced with permission.[163 ] Copyright
2023, Elsevier.
exhibited enhanced high-voltage stability and environmen-
tal/thermal stability.[] To date, little research has been con-
ducted on Na-site substitution. However, other potential dopants
such as Mg, Zn, Zr, Li, K, and Bi provide the possibility of fur-
ther extending the cycling life of layered-oxide cathodes.[– ] It
should be noted that the substitution on Na sites may result in
more initial capacity loss than the substitution of TMOlayers,
and the narrowed ion-diusion channel adversely aects the re-
action kinetics.[, ] Therefore, it is necessary to carefully select
appropriate doping agents and adjust their concentrations to bal-
ance the trade-os between cycling life, capacity, and rate capabil-
ity. In addition to cation substitution, the combination of Fions
with strong electronegativity to enhance the stability of lattice-O
has also proven eective in improving the electrochemical per-
formance of layered oxides.[ ] It was demonstrated that the in-
troduction of F ions reduces the thickness of the TMOlayer,
but broadens the ion-diusion channel to improve reaction kinet-
ics, alleviates abrupt phase transitions to suppress the formation
of microcracks, and enhances the resistance to electrolyte cor-
rosion to alleviate the dissolution of active materials.[ ] There-
fore, pouch-type full cells assembled with optimized cathodes
(Na.[Ni.Mn. ]O.F.) and HC anodes exhibited excellent
cycling stability, with a capacity retention of % after  cycles
at  mA g. However, it should be noted that the introduction
of toxic F ions is unpopular in the industry, which limits the prac-
tical potential of these materials.
The frequent phase evolution accompanied by sluggish re-
action kinetics and multi-step redox characteristics is responsi-
ble for the poor electrochemical performance of NaNi.Mn.O.
Therefore, it is recommended to optimize the composition
to alleviate high-pressure phase transformation, thereby sup-
pressing the formation of microcracks and avoiding mechanical
failure.[ ] A systematic investigation of the eects of Sn doping
on the electrochemical performance and structural integrity of
NaNi. Mn.Oshowed that the introduction of Sn can alleviate
the abrupt contraction of crystal-cell parameters caused by the
P’ to O’ phase transition at high voltage (Figure 6a), thereby
inhibiting the formation of microcracks and extending the cy-
cling life; the capacity retention after  cycles at . C increased
from . to .% after Sn addition.[ ] In addition, Ti+, with
a similar ionic radius but a greatly dierent redox potential from
Mn+, was shown to improve the TM–O bonding energy, increase
the working potential, and disrupt Na/vacancy ordering.[] An
Mg/Ti co-substitution strategy was proposed to improve the
structural stability of NaNi.Mn.Ocathodes, in which Ti ex-
pands the interlayer spacing, and Mg acts as an HF scavenger.[]
Adv. Energy Mater. 2023, 2301975 © 2023 Wiley-VCH GmbH
2301975 (14 of 41)
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202301975 by Hanyang University Library, Wiley Online Library on [18/08/2023]. 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
www.advancedsciencenews.com www.advenergymat.de
Unlike the drastic phase transition from P to O’ in non-
substituted samples under high voltage, co-doped samples with
extended TM–TM distances and suppressed Na/vacancy order-
ing provide a smooth phase transition, delay the O–O’ phase
transition, and form a stable PO intermediate buer phase in-
stead of the O’ phase (Figure b). Therefore, the optimized cath-
ode (Na[Ni/Mn/Mg/ Ti/ ]O) exhibits excellent structural
integrity as it eectively suppresses microcracks caused by the
significant shrinkage of the interlayer spacing. A systematic in-
vestigation of the structure and electrochemical performance of
Ti- and Cu-doped NaNi.Mn.Oshowed that the crystal-cell pa-
rameters and oxidation peak potential of the samples increased
with increasing Ti/Cu substitution concentration (but the solu-
bility of Cu was limited to below . mol).[ ] The co-doped sam-
ple formed a mixed phase in the high-voltage region instead of a
new O phase, thereby reducing the unit-cell volume shrinkage.
Therefore, the optimized cathodes (NaNi.Cu.Mn. Ti. Oand
NaNi. Cu.Mn. Ti. O) exhibited a higher energy density
than the reference NaV(PO)Fmaterial. In addition, it was
demonstrated that co-substitution of Ti and Zn/Mg/Cu can also
suppress high-voltage phase transitions and reduce unit-cell vol-
ume changes, thus eectively improving the cycling stability of
NaNi. Mn.O.[] However, due to the presence of O domains
related to cation migration at the fully charged state, the cycling
stability of the modified cathodes is still unsatisfactory.
For the above reasons, by adjusting the electronic structure of
the TMOslab to expand the solid-solution zone, the structural
reversibility and Na-ion diusion ability of the obtained cathode
is fundamentally improved.[ ] For instance, Sn+(d) with
filled d orbitals cannot interact with O p orbitals, so orbital hy-
bridization occurs between the remaining TM and O atoms, thus
preventing charge delocalization in the TMO-layer, increasing
the output voltage, and improving the structural reversibility at
high voltages.[ ] The O–P phase transition can be delayed,
and the redox potential can be increased by the chemical substitu-
tion of Mn in NaNi.Mn.Oby Sn (Figure c).[] The improved
cycling stability of the resulting cathode (NaNi.Sn.O)comes
at the cost of sacrificing some energy density due to the introduc-
tion of heavier Sn. In addition, Ti-doping can eectively alleviate
multi-step phase transformation because the slightly larger vol-
ume of TiOoctahedra than MnOoctahedra enables flexibility
to absorb structural distortion and strain caused by Na-ion in-
sertion/extraction processes.[] The introduction of Ti+with a
large ionic radius broadens the TM–O bond length, inhibiting
the phase transition above . V to broaden the solid solution re-
action zone of the P phase (three charging/discharging plateaus
become sloped curves) (Figure d).[,] Therefore, the opti-
mized cathode NaNi.Mn. Ti. Oexhibited long cycling stabil-
ity (% capacity retention after  cycles at C) and fast ionic
diusion (. × cms) over the voltage window of .–
. V.[ ] More importantly, the air stability of layered oxides
can be eectively improved by introducing dopants with large
ionic radii and dierent Fermi levels to weaken the hybridiza-
tion between O(p) and TM orbitals and enhance the Na–O bind-
ing energy (reducing the Na layer spacing).[,] Therefore,
NaNi. Mn.Ti.O[ ] and NaNi.Cu.Mn. Ti. O[ ] ex-
hibit excellent air stability and do not undergo structural changes,
even when soaked in water. It should be noted that due to the
structural similarity between the O- and P-type phases, some
studies mistakenly attribute the O-phase structure in the high-
voltage region to the P-type phase, thereby misidentifying the
solid-solution reaction zone.[, ] For O and P phases, the
intensity ratio between () and () diraction peaks can be
used as an indicator of the crystal structure, where the O and P
phases exhibit higher relative intensities of () and () peaks,
respectively.[]
3.2.3. NaTM1TM2O2
Owing to their abundant natural raw materials and low environ-
mental impact, Fe/Mn-based cathode materials are also attractive
systems. O-Na.Mn/Fe/Odelivered a discharge capacity of
 mA h gover a voltage window of .–. V, but suers from
poor cycling stability due to the complex phase transitions dur-
ing the charge/discharge process.[] Although cation doping ef-
fectively improves the electrochemical behavior of Fe/Mn-based
cathodes, the abnormal ICE caused by Mn+/Mn+/Mn+redox
reactions at low voltage limits their commercial application.[]
Based on the typical O–P phase transition, NaNi.Fe. Ode-
livered a reversible capacity of  mA h gand exhibited a
smooth charging/discharging profile and stable cycling stability
in the voltage range of .–. V.[ ] However, the rapid capac-
ity degradation caused by Fe migration and irreversible changes
in the lattice parameters under high voltage still need to be ad-
dressed. Owing to the high price of Co, the large-scale application
of Co-containing cathode materials will weaken the competitive-
ness of SIBs. Furthermore, unlike LIB systems, the introduction
of Co may lead to cation mixing and weaken the layered structural
characteristics,[ ] making Co-containing layered oxides unsuit-
able as commercial cathode materials for SIBs. Moreover, cath-
ode materials with cationic-ordered superlattice structures (such
as honeycomb structures) can eectively inhibit the intercala-
tion reaction of HO due to enhanced interlayer interactions,
which enhances the air stability of layered oxides.[ ] A novel O’-
NaMn. Al.Ocathode with a superlattice layered structure was
obtained by Al substitution, in which AlOand MnOoctahe-
drons show a “queue-ordered” arrangement in TMOslabs.[]
The ordered arrangement of TM ions suppresses the lattice dis-
tortion caused by the JT eect, while the introduction of Al ions
induces electron transfer to enhance the TM–O bonding strength
and reduce the formation energy of layered structures. Therefore,
the capacity retention rate of the optimized cathode increased
from . to .% after  cycles at  mA g. In addition,
some high-voltage binary metal-oxide cathode materials have
been presented. For example, a novel high-voltage cathode, O-
Na.Ni.Sn. O, with a . V operating voltage, demonstrated
a capacity of  mA h gdue to the contribution of the Ni+/Ni+
redox couple within a voltage window of .–. V.[ ] However,
subsequent studies showed that this material completely trans-
formed into its hydrated phase after being exposed to humid air
for h, indicating its strong sensitivity to air.[] In recent years,
honeycomb layered oxides (NaNi/M/O, where M represents
As, Sb, and Bi) with high operating voltage have also attracted the
attention of researchers, as the small size of M enables more dis-
tortion of the NiOoctahedron and a decrease in the energy level
of the egorbitals.[ ] However, the presence of excessive inactive
ions leads to low reversible capacity (about  mA h g)and
Adv. Energy Mater. 2023, 2301975 © 2023 Wiley-VCH GmbH
2301975 (15 of 41)
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202301975 by Hanyang University Library, Wiley Online Library on [18/08/2023]. 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
www.advancedsciencenews.com www.advenergymat.de
Figure 7. a) Schematic diagram of the relationship between the discharge capacity, capacity retention, and thermal stability of O3-Na[NixCoyMnz]O2
cathodes. Reproduced with permission.[178 ] Copyright 2016, the Royal Society of Chemistry. b) Cycling performance of Na[Ni0.75xFexMn0.25 ]O2at 0.2
and 0.5 C rate over 1.5–3.9 V. Reproduced with permission.[ 183 ] Copyright 2016, Elsevier. c) Relationship between the voltage and capacity of Nax(Cu–
Fe–Mn)O2systems, d) as well as their corresponding cycling performance over 1.6–2.2 V. Reproduced with permission.[184] Copyright 2020, Elsevier.
increased production costs, which limits its potential for large-
scale commercialization.
3.2.4. NaM1M2M3O2
Inspired by the successful commercialization of NCM cathodes
in LIBs, researchers systematically studied the electrochemical
and thermodynamic behavior of O-Na[NixCoyMnz]O(/ x
.) cathodes.[ ] Similar to lithium-based systems, increasing
the Ni content, especially above %, may increase the capacity,
but at the expense of the cycling life, rate properties, and ther-
mal stability (Figure 7a). This confirms the importance of balanc-
ing the TM composition of layered oxides. NaNi/Mn/Co/O
with relatively high thermal stability and cyclability undergoe a
multi-step phase transition of O–O–P–P during the sodia-
tion/desodiation process, where the ab plane slightly contracts
and the interlayer spacing continues to expand.[ ] Ti-doped
NaNi. Co.Mn.Oshowed enhanced cycling stability (capac-
ity retention increased from  to % after  cycles at . C)
and thermal stability (exothermic peak increased by  °C) with-
out sacrificing reversible capacity ( mA h g).[ ] In addition,
this Ti-substitution strategy has been proven to improve air sta-
bility; the original crystal structure was maintained even when
immersed in water, with only a slight decrease in crystallinity.[ ]
It should be noted that unlike Li-ion systems, the development of
high-Ni layered oxides in SIB systems does not seem to be a pop-
ular research direction because Ni triggers the multiphase trans-
formation and lattice defects, and is quite expensive.[ ] There-
fore, surface engineering and optimizing the chemical composi-
tion and morphology are the key directions in the development
of Ni/Co/Mn-based cathodes.
The Ni/Fe/Mn-based cathode materials derived from Ni/Mn-
binary metal oxides have also received attention from re-
searchers. The introduction of Fe with a large ionic radius ex-
pands the TMOlayer and facilitates electronic delocalization.
Therefore, NaFe.Ni.Mn. Oshowed a reversible Ohex
Phex phase transition over a voltage range of .–. V, without
forming other monoclinic structures, thus delivering higher aver-
age operating voltage (. V versus . V for NaNi.Mn.O),
reversible capacity (. mA h g), and cycling stability (%
capacity retention after  cycles at C).[ ] However, it was
proposed that NaFe/Ni/Mn/ Otransforms into a monoclinic
O’ phase at . V and undergoes a multi-step phase transition
of O’–P’–O during discharge.[ ] In addition to the chemi-
cal composition, the proportion of each component in the TMO
slabs aects the electrochemical performance of the active mate-
rial. It was demonstrated that the Fe+/Fe+redox couple is quite
reversible at low Fe concentration (/ per unit), but significantly
irreversible at high content (. per unit).[ ] Therefore, opti-
mizing the composition and proportion of active elements is the
key to the commercialization of SIBs. A systematic investigation
of the electrochemical performance of O-NaNixFeyMnzOwith
dierent compositional ratios showed that NaNi.Fe. Mn.O
exhibited the best cycling stability due to the suppressed JT ef-
fect and expanded interlayer spacing.[ ] However, the redox cou-
ple of Mn+/Mn+activated at low voltage decreases the energy
density. Na[Ni. Fe. Mn.]Owith high Ni content delivered
a high discharge capacity of  mA h gand an average oper-
ating voltage of . V within a voltage window of .–. V.[]
Adv. Energy Mater. 2023, 2301975 © 2023 Wiley-VCH GmbH
2301975 (16 of 41)
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202301975 by Hanyang University Library, Wiley Online Library on [18/08/2023]. 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
www.advancedsciencenews.com www.advenergymat.de
However, it undergoes a complex phase transition from O to
O’–P–O’’ during the desodiation process, resulting in infe-
rior ion-diusion kinetics and the formation of microcracks. Al-
though limiting the charging cuto voltage to . V ensures that
the developed cathode maintains the P phase structure after
the extraction of . mol of Na, this sacrifices part of the re-
versible capacity. It was demonstrated that an appropriate in-
crease in the Fe content can alleviate unstable Ni+-induced O-
loss, thereby eectively improving the cycling and thermal sta-
bility of Ni/Fe/Mn-based cathode materials (Figure b).[ ] In
addition, cation substitution (such as Al, Ti, Zn, Mg, and Li) was
shown to further improve the high-voltage stability.[,] For ex-
ample, due to the introduction of Al enhancing the TM–O bonds,
widening the ion-diusion channels, alleviating the JT eect of
Mn+, and reducing the interfacial resistance, the resulting cath-
ode (NaNi/Mn. Fe/ Al. O) exhibited improved rate capa-
bility (. mA h gat C) and cycling stability (% capacity
retention after  cycles at C).[ ] An investigation of the struc-
ture and electrochemical performance of NaMn.Ni. Fe. O
and NaMn.Ni. Fe. Mg.Ocathode materials demonstrated
that Mg doping reduces the JT eect through a charge compen-
sation mechanism and improves the ionic diusion capability
by expanding the interlayer spacing.[ ] Therefore, the capacity
retention of Mg-doped samples after  cycles at . C signifi-
cantly increased from  to %. The incorporation of Ti+with-
out d electrons strengthens the binding energy between TM ions
and O by increasing the O-electron density while increasing the
electrostatic repulsion of the O slab to expand the interlayer dis-
tance and reduce the Na-ion migration energy barrier.[  ] Hence,
Na[(Mn.Fe.Ni.). Ti. ]Oexhibited improved cycling stabil-
ity, with a capacity retention of % after  cycles at . C,
accompanied by some reduction in the initial discharge capac-
ity (from  to  mA h g).[] Furthermore, it was demon-
strated that the introduced Ti ions are enriched on the surface
of bulk materials, which eectively suppresses monoclinic dis-
tortions and reduces the reactivity with water, thereby improving
the cycling and water stabilities.[] The morphology and crystal
structure of Na.Ni. Fe. Mn.Ti.Oremained unchanged
after days of exposure to humid air and h of immersion in
water, exhibiting high air stability (only .% reversible capacity
loss after days of exposure) and excellent cycling stability (.%
capacity retention after  cycles at C).[ ] It was shown that
doping with Li (. mol%) to support the TMOlayer can sup-
press the migration of Fe+ions, thereby improving the reversible
capacity (. mA h gwithin .–. V), cycling stability
(.% capacity retention after  cycles at . C), rate capability
(. mA h gat C), and thermal stability ( °C exothermic
peak) of Na[Ni.Fe. Mn.]O.[ ] Furthermore, it was demon-
strated that the introduction of Li not only strengthens the TM–
O bonds and reduces JT distortions to improve the structural
reversibility under high potentials. Moreover, it also acts as a
sacrificial agent to react with Fions to form a stable cathode–
electrolyte interphase (CEI) layer containing LiF to alleviate the
dissolution of the active material.[ ] Therefore, the optimized
cathode (O-Na.Li.Ni.Mn. Fe. O) maintained % of
its initial capacity after  cycles over a wide voltage range of .
to . V at C.
Because Cu plays an active role in improving the structural
and air stabilities of layered oxides, Fe/Mn/Cu-based cathode ma-
terials composed of inexpensive metal ions are also considered
promising materials.[] The O-Na.[Cu.Fe. Mn.]Ocath-
ode with Cu/Fe active centers undergoes multiple phase transi-
tions from O to P and finally to O’ during the desodiation
process (.–. V), providing a reversible capacity of  mA
hg
and an average storage voltage of . V.[ ] It was pro-
posed that the introduction of Cu reduces the average valence
state of Mn ions (forming electrochemical inactive Mn+)toavoid
the P–O phase transition and improve the cycling stability, while
the Cu+/Cu+redox couple compensates for the partial capac-
ity loss.[ ] In addition, a correlation between the solubility of
Cu and the valence state of Mn was demonstrated. An aver-
age oxidation state of Mn above +. induces the segregation
of CuO to satisfy a new electrostatic equilibrium that accommo-
dates low-valent Mn. Furthermore, the cation composition dia-
gram of the Nax(Cu–Fe–Mn)Osystem was investigated, and it
was shown that low-valent Mn provides high reversible capacity
(for example,  mA h gfor Na.(Cu.Fe. Mn. )O), Fe re-
duces the average operating voltage, and Cu improves the rate
performance, cycling stability, and air stability (Figures c,d).[]
In addition to the composition of the TMOlayer, the eect of
the Na content on electrochemical performance was studied.[]
It was demonstrated that reducing the Na content in the O-
NaxCu.Fe.Mn. Ocathode to a critical value (.) of the
O/P dual-phase structure can promote the phase evolution
of O–P, avoid changes in the unit-cell parameters of the O
structure and minimize voltage hysteresis. It should be noted
that similar to Fe/Mn-based cathode materials, the Mn+/+/+
redox reactions delivers high reversible capacity but low operat-
ing voltage, which reduces the energy density and leads to an ab-
normal ICE.[, ] Therefore, the Fe/Mn/Cu-based cathodes are
typically cycled over a narrow voltage window (for example, .–
. V), thereby exhibiting good cycling stability (% capacity re-
tention after  cycles at C) but low reversible capacity ( mA
hg
).[ ] A systematic investigation of the eect of Li-doping
on the electrochemical performance of NaCu.Fe. Mn. O
showed an inhibition of the JT eect and enhanced Fe+/+reduc-
tion, resulting in an optimized cathode with high reversible ca-
pacity, average discharge voltage, and cycling stability.[ ] It was
further demonstrated that the surface degradation caused by Mn
reduction and side reactions with electrolytes further degrades
the cycling stability of layered oxides, which implies that further
optimization of the surface structure is required.[]
3.2.5. High-Entropy Compounds
In recent years, high-entropy oxides with high fracture tough-
ness, tolerance to a wide operating temperature range, high
electronic delocalization, and high structural stability have re-
ceived extensive attention in the fields of electrochemical en-
ergy storage and electrocatalysis.[ ] Although a precise defini-
tion of high-entropy oxides is still controversial, they are usually
defined based on composition (single-phase complexes contain-
ing at least five metal atoms, with concentrations ranging from
to % per atom) or configurational entropy (configuration en-
tropy greater than .Rat room temperature, independent of
the phase number).[ ] Due to the synergistic eect of multiple
metal elements, high-entropy oxides with cocktail eects exhibit
Adv. Energy Mater. 2023, 2301975 © 2023 Wiley-VCH GmbH
2301975 (17 of 41)
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202301975 by Hanyang University Library, Wiley Online Library on [18/08/2023]. 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
www.advancedsciencenews.com www.advenergymat.de
Figure 8. a) In-situ XRD patterns of high-entropy oxides in the first two cycles within the voltage range of 2.0–3.9 V. Reproduced with permission.[ 201 ]
Copyright 2020, Wiley-VCH GmbH. b) Schematic diagram of the structural evolution of the high-entropy configuration and conventional cathodes
after repeated cycling. Reproduced with permission.[207 ] Copyright 2022, American Chemical Society. c) Schematic diagram of the phase transitions of
cathode material with high-entropy and superlattice structures during charging/discharging. Reproduced with permission.[209] Copyright 2022, Wiley-
VCH GmbH. d) Charging/discharging profile, cyclic voltammogram, and long-term cycling stability of high-entropy substituted cathodes. Reproduced
with permission.[210 ] Copyright 2023, Elsevier.
surprising electrochemical properties. As early as , a novel
quinary layered oxide cathode O-NaNi/Co/Fe/ Mn/Ti/O
was presented.[ ] The K-edge XANES results indicated that Ni,
Fe, and Co are responsible for charge compensation upon sodi-
ation/desodiation, the introduction of Ti enhances the ionic dif-
fusion ability, and the addition of Mn improves charge-transfer
properties. Similarly, the charge compensation reaction of the
five-component cathode Na.Ni/ Fe/ Co/ Mn/Ti/Oduring
charging/discharging is mainly attributed to the Ni+/Ni+re-
dox pair, with some contributions from Fe+/Fe+and Co+/Co+
pairs, while Ti+and Mn+act as structural pillars.[] Because
the multi-component structure suppresses unfavorable phase
transitions, the obtained cathode material only exhibited a voltage
plateau attributed to O and P phase transitions at ./. V,
along with a sloping voltage curve related to the solid solu-
tion reaction of the P structure between . and . V. This
extended solid-solution process and negligible unit-cell volume
change (.%) endowed the resulting five-component cath-
ode material with excellent structural reversibility and cycling
stability.[] The O-type high-entropy oxide cathode material
NaNi. Cu.Mg. Fe. Co. Mn.Ti.Sn. Sb.Owas first
presented in , where Ni, Cu, Co, and Fe provide charge com-
pensation to achieve a high reversible capacity, Mg and Ti acts
as pillars to maintain the layered framework structure, Mn di-
rects the structure, and Sn and Sb increase the average operat-
ing voltage.[ ] The redox elements in small-component O-type
cathodes are uniformly distributed in the TMOslab, facilitating
the phase transition from O to P during charging/discharging.
However, the local changes in the redox elements in this multi-
component high-entropy cathode are easily accommodated by
the local characteristics of other stabilizers, thereby expanding
the region of the O phase structure (% of total capacity) and
increasing the energy density (average operating voltage of . V)
(Figure 8a). This delayed phase transition was also observed
in high-entropy O-type NaCu.Ni. Fe. Mn.Ti.Ocathodes
that maintained the O-type structure until . Na ions were ex-
tracted from the lattice unit.[ ] Although the high-entropy cath-
ode exhibited excellent cycling stability, with a high capacity re-
tention of % after  cycles at C, a high proportion of inac-
tive elements resulted in a low reversible capacity of  mA
hg
. To balance high entropy and high capacity, they further
optimized the concentration of each component in the TMO
slabs (NaNi. Mg.Cu.Fe.Mn. Ti. Sn.O).[ ] This high-
entropy configuration enhances the surface energy of the ()
plane to increase the Na-ion transport channel, while the TMO
slab undergoes slight bending rather than producing a rock-salt
phase near the surface to accommodate the volume changes of
the unit cell during charging/discharging (Figure b). Therefore,
the obtained cathode exhibited high reversible capacity (. mA
hg
), excellent structural integrity (less TM dissolution and
cracks), long cycling life (% capacity retention after  cycles
at C), and thermal stability (exothermal peaks at  °C) in a nar-
row voltage window of . to . V. A further study investigated
the electrochemical properties of NaMn.Fe. Co. Ni. Ti. O
and Na/Li/Fe/Co/Ni/Mn/ Oand found that the quick
phase transformation from O to P metallic phase is the key to
Adv. Energy Mater. 2023, 2301975 © 2023 Wiley-VCH GmbH
2301975 (18 of 41)
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202301975 by Hanyang University Library, Wiley Online Library on [18/08/2023]. 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
www.advancedsciencenews.com www.advenergymat.de
Figure 9. a) Energy cascade and physical model of the formation of P2-Na2/3CoO2. Reproduced with permission.[212 ] Copyright 2020, Springer Nature.
b) Phase evolution of Na0.7Ni0.2 Cu0.1Fe0.2Mn0.5 O2–𝛿monitored by high-temperature in-situ XRD. Reproduced with permission.[ 216] Copyright 2022,
Elsevier. c) High-resolution TEM (HRTEM) image of a P2/O3 biphasic cathode material. Reproduced with permission.[218] Copyright 2014, Wiley-VCH
GmbH. d) Relationship between the content of P2 phase structure and radial potential. Reproduced with permission.[219 ] Copyright 2023 Springer Nature.
e) Schematic diagram of the structural evolution of P2, O3, and P2/O3 materials during charging. Reproduced with permission.[220 ] Copyright 2022,
Elsevier. f) Comparison of the overall performance of three different structured cathodes. Reproduced with permission.[221 ] Copyright 2022, Wiley-VCH
GmbH.
the high Na-ion diusion kinetics (Figure c).[,] It is neces-
sary to further clarify the optimal compositions and phase transi-
tion processes of high-entropy oxides. Recently, a novel layered
cathode material, NaNi.Mn.Co.Cu. Fe. Li.Ti.Sn. O,
was developed via high-entropy substitution, where Ni, Co, and
Mn provide charge compensation, Fe, Cu, and Ti serve as sta-
bilizers, Li alleviates lattice distortion, and Sn increases the
working potential.[ ] This multi-component high-entropy oxide
smoothed the successive phase evolution of O O’ P,
thereby delaying the phase transition from P to P’, and increas-
ing the Young’s modulus and indentation hardness resulting in
a significant improvement in the cycling performance of the re-
sulting cathode (.% capacity retention after  cycles at
 mA g) (Figure d). Although the introduction of inactive el-
ements slightly reduces the reversible capacity (. mA h g),
the higher average discharge voltage of this high-entropy cathode
(. V versus . V for NaNi.Mn.O) provided a higher energy
density (. mW h g). In addition, aging tests showed that
high-entropy structures can eectively alleviate phase changes,
TM dissolution, and the formation of NaCOimpurities. Based
on the discussion above reasons, high-entropy oxides are cur-
rently in the early research stage, but are promising candidates
for cathode materials in advanced SIBs. Future research needs
to focus on high capacity, high-voltage performance, air stability,
low cost, and morphology/interface engineering.
3.2.6. Multiphase Structure
The crystal structure of Na–TM–O compounds is aected by ther-
modynamics (such as the Na/TM ratio and synthesis tempera-
ture) and kinetics (including the preparation method, raw ma-
terials, and synthesis conditions), so they show rich structural
diversity.[] A systematic investigation of the phase evolution
process of Na.TMO(TM =Co, Mn) suggested that the for-
mation of the first phase is controlled by thermodynamics, while
subsequent transitions are related to kinetics (Figure 9a).[]
Considering the structural diversity of layered oxides, it is
reasonable to consider the development of multiphase cath-
ode materials with dierent electrochemical properties, which
exhibit better performance than single-phase materials due
to synergistic eects.[ ] For example, the zNa/Ni/ Mn/O
(z)NaNi.Mn. O(/ x) system could have a two-
phase structure, consistent with the experimental results of
P-Na.Ni. Mn.Oand O-NaNi. Mn.O.[ ,] Aseriesof
cathodes (NaxNi.Mn. O;x=., ., and .) were pre-
pared with a mixed P/O structure by changing the sodium
content.[ ] However, due to the charge compensation mecha-
nism, the presence of Na vacancies requires an increase in the
content of JT-active Ni+and even the formation of NiO impu-
rities, which may have adverse eects on the electrochemical
performance. The synthesis phase diagram of NaxMnyNiyO
materials showed that proper chemical compositional design is
the key to the synthesis of two-phase structures.[ ] A doping
method was used to control the ratio of the two-phase struc-
ture, demonstrating that the dopant-induced P phase forma-
tion trend is Mn >Ni >Cu >Fe >Co >Ti. In addition, by
controlling the sintering temperature, cooling rate, and calci-
nation time, even cathodes with the same chemical composi-
tion can form dierent crystal structures.[, ] Through theo-
retical simulations and XRD analysis, it was demonstrated that
the formation of the multiphase structure (P/O) originated
from fluctuations in the local cation potential due to the un-
even distribution of metal ions, which is mainly dependent on
Adv. Energy Mater. 2023, 2301975 © 2023 Wiley-VCH GmbH
2301975 (19 of 41)
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202301975 by Hanyang University Library, Wiley Online Library on [18/08/2023]. 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
www.advancedsciencenews.com www.advenergymat.de
temperature-dependent reaction kinetics and thermodynamics
(Figure b).[] Similarly, a NaxNi/Co/ Mn/Ocathode with a
three-phase structure (intergrowth P/O/O structure) was suc-
cessfully synthesized by controlling the calcination temperature
and cooling step (quenching).[ ] Unlike the specific synthesis
conditions of single-phase cathode materials, the optimal synthe-
sis conditions and reaction mechanism of multiphase materials
still need further clarification.
To combine the advantages of P-type and O-phase cathodes,
a series of layered cathode materials with two-phase or three-
phase structures have been developed. Previous studies have
demonstrated that the incorporation of heterometals (especially
Li) into TMOlayers can facilitate the formation of mixed-phase
structures, as these dopants can oer dramatically dierent elec-
tronic configurations, oxidation states, and sizes.[,, ] John-
son et al. first incorporated Li as a dopant into NaNi.Mn.O
to induce the formation of a P/O biphasic cathode mate-
rial (Na.Li.Ni.Mn. O+d) with a nanoscale topotactic inter-
growth structure (Figure c).[] The volume change of the P
domain during the sodiation/desodiation process is less than
%, while the O phase transforms into the P phase and re-
mains unchanged throughout the discharge process, which en-
ables the presented cathode to exhibit improved structural sta-
bility and high-rate capability. However, this O-majority ma-
terial still suers from low cycling lifetimes, which needs to
be resolved. Later, a novel biphasic cathode material was de-
veloped, (Na.Li.Mn.Ni. Co. O+𝛿), which integrates a
small amount of O into a P-dominated layered material
to meet the requirements of high capacity and excellent cy-
cling stability.[] The developed cathode exhibits a high re-
versible capacity ( mA h g) and good cycling stability (%
capacity retention after  cycles). Recently, a novel penta-
nary metal-oxide cathode with a P/O biphasic structure was
presented, namely Na/Ni/ Mn/Fe/Mg/ Li/O.[] Because
multi-element doping inhibits the dissolution of active materi-
als, the structural reversibility was improved. Furthermore, the
topologically intergrown structure suppressed the high-voltage
phase transition and lattice mismatch, resulting in a cathode with
excellent rate properties (. mA h gat  C) and cycling
stability; the capacity retention increased from  to % after
 cycles at C. It should be noted that Li-substituted cathode
materials do not all exhibit a layered structure, and the intro-
duction of excess Li can result in the formation of anomalous
spinel phases. The Na.Li.Ni.Fe.Mn. O+𝛿cathode mate-
rial with a Li-substituted layered (O) and spinel intergrowth
structure was prepared with an alkali metal content greater than
.[ ] Compared with the pristine samples, the layered/spinel in-
tergrown cathode exhibited high rate capability ( mA h gat
 mA g) and Na-ion diusion coecient (. × cms)
because the additional spinel phase provides D ion-diusion
channels to facilitate ion transport.
Although the P/O heterostructure induced by Li doping re-
sults in excellent Na-ion storage performance, the introduction of
Li leads to an increase in the production cost of the cathode ma-
terial, which reduces the scalability of SIBs. Therefore, cations
(such as Ti, Sn, Sb, Mg, Cu, and Fe) with dierent ionic radii and
valence states are introduced to induce the formation of multi-
phase structures.[– ] Hu et al. proposed that the crystal type of
the Na layered oxide can be determined from the cationic poten-
tial, which is beneficial for guiding the design of O- and P-type
cathodes.[ ] However, this empirical formula has limited impact
on the development of non-equilibrium and distorted phases.
The influence of chemical composition changes on the propor-
tion of the bi-phasic structure was evaluated using a radial poten-
tial predictor (combining the cation potential and ionic radius)
(Figure d).[] The proportion of P phase in the cathode ma-
terials with the same Na content decreased linearly with increas-
ing radial potential. Structural analysis and theoretical calcula-
tions showed that the intergrowth structure with an interlocking
eect can eectively inhibit unfavorable cation migration and O-
loss, thus improving the structural stability and reversibility of
the anion redox reaction. In addition, the local chemistry modu-
lation of dopants can induce the formation of P/O intergrowth
structures by regulating the structural evolution and formation
energy, thereby suppressing the sliding of TMOslabs and en-
hancing the reversibility of crystal structures.[ ] Hence, Ti with
low cationic potential was introduced (Na.Ni. Mn.Ti.O)to
trigger the formation of an O structure and adjust its pro-
portion in the biphasic cathode.[ ] Benefiting from the sup-
pressed high-voltage phase transition and reduced lattice param-
eter changes, the developed heterostructured cathode exhibited
improved structural integrity and excellent cycling stability (with
a capacity retention of % after  cycles at C and a reversible
capacity of  mA h gat . C within .–. V). Moreover,
an interfacial interlocking reaction mechanism was proposed to
explain the excellent Na-ion storage performance of heterostruc-
tured cathodes, whereby the mutual anchoring of dierent crys-
tal domains reduces the strain energy and mechanical damage at
the phase boundary (Figure e).[] Accordingly, the optimized
P/O-Na.Ni. Mn.Ti.Ocathode exhibited high cycling
stability with a retention rate of .% after  cycles at C.
Because the introduction of dopants can reduce the capac-
ity, the strategy of synthesizing multiphase structures by ad-
justing the sintering temperature and/or Na content is receiv-
ing increasing attention. In addition, introducing Na vacancies
to reduce the amount of surface residual Na and improve the
reaction kinetics improves the Na-ion storage performance of
O-type cathodes, making it possible to achieve long cycling
life under high reversible capacity.[,] A series of  com-
pounds were synthesized by adjusting the Na content and cal-
cination temperature to evaluate the temperature dependence
of the cathode structure.[ ] Although the in-situ XRD results
suggested that the optimized biphasic cathode undergoes com-
plex structural evolution during the sodiation/desodiation pro-
cess, it was proposed that the interlocked intergrowth struc-
ture can reduce internal stresses and alleviate lattice parame-
ter changes (biphasic clamping reaction mechanism), leading
to improved electrochemical performance (Figure f). A three-
phase cathode (P/P/O-Na.Ni. Mn.O) was synthe-
sized by only adjusting the calcination conditions (temperature
and atmosphere).[ ] Benefiting from the structural constraint
eect of this multiphase structure that suppresses detrimen-
tal phase transitions, the developed cathode material showed
high rate capability ( mA h gat  mA g) and cy-
cling stability (the residual capacity increased by  mA h g
after  cycles). A series of P/O mixed-phase intergrowth
layered cathode materials were synthesized by optimizing
the Na content (combining P-Na.[Ni.Fe. Mn. ]Oand
Adv. Energy Mater. 2023, 2301975 © 2023 Wiley-VCH GmbH
2301975 (20 of 41)
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202301975 by Hanyang University Library, Wiley Online Library on [18/08/2023]. 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
www.advancedsciencenews.com www.advenergymat.de
O-Na[Ni.Fe. Mn.]O).[ ] In-situ XRD results showed that
the optimized biphasic cathode undergoes a reversible phase
transition and negligible volume change during cycling, thereby
providing high reversible capacity (. mA h g) and improved
cycling stability (% capacity retention after  cycles at .C)
over a wide voltage window of .–. V.
3.3. Potential Coating Medium for O3-Type NaTMO2
To suppress interfacial parasitic reactions and alleviate the disso-
lution of active materials, researchers proposed a surface coating
strategy to improve the interface stability, including coatings of
C,[, ] oxides,[, ] fluorides,[ ] and phosphates.[ ,] The
introduced surface-modification layers have the following func-
tions: ) Improve ionic/electronic conductivity. A surface modifi-
cation layer with high ionic/electronic conductivity can eectively
improve the surface properties of cathode materials and reduce
the internal reaction resistance.[] ) Physical protection. The
coating isolates the active material from the corrosive electrolyte
to prevent the corrosion by harmful by-products such as HF. In
addition, the coating also acts as a buer to alleviate the volume
change of the crystal during the charging/discharging process,
thus improving the integrity of the composite. ) Reduce resid-
ual Na on the surface. As mentioned in section , residual Na on
the surface may cause gel of the slurry and loss of capacity.[]
Therefore, in-situ construction of the Na-containing coating can
eectively improve the electrochemical performance and inter-
facial stability by consuming the insulating surface residue.[ ]
It should be mentioned that owing to the sensitivity of Na-rich
compounds to humid air, the post-treatment of sodiated TM ox-
ides may lead to surface degradation.[] Therefore, it is necessary
to carefully explore the synthesis process and preparation envi-
ronment.
3.3.1. Carbon Coating
Carbon-based materials with high electronic conductivity and an
elastic network are eective as a coating layer, so they are widely
used to modify electrode materials for LIBs and SIBs. However,
most Na–TM–O composites are calcined in air or O,sothisC-
coating strategy is usually applied to specific cathode materials,
such as Cr- and V-based oxides.[] Carbon-coated Fe/Mn-based
cathode materials can be achieved using solid-state-assisted pro-
cesses or secondary high-temperature carbonization steps (Ar
atmosphere).[, ] However, due to unstable coatings or intro-
duced O-vacancies, the resulting cathode materials exhibit unsat-
isfactory electrochemical performance. Therefore, from the per-
spective of production costs and Na-ion storage performance, the
C-coating strategy does not seem to be fully suitable for modify-
ing layered oxide materials.
3.3.2. Oxide Coatings
In addition to C-coatings, metal oxides are widely used for surface
modification of layered oxide cathodes owing to their good con-
ductivity, structural stability, and interfacial compatibility.[,]
For instance, atomic layer deposition technology was used to
precisely encapsulate NaNi.Mn.Oin an AlOcoating with
controllable thickness ( nm) to reduce side reactions and sup-
press electrolyte corrosion (Figure 10a).[] Furthermore, the sur-
face of O-Na[Ni.Co.Mn.]Ocathodes were uniformly coated
with nano-AlOvia dry ball-milling in a dry room to prevent the
formation of surface by-products.[ ] The introduced protective
coating can scavenge the HF produced by the adverse parasitic re-
action with the electrolyte, accompanied by the formation an AlF
layer on the outer surface, thus inhibiting the dissolution of TM
ions and the generation of microcracks (Figure b). Therefore,
the pouch-type full-cell assembled with this surface-optimized
cathode and HC anode exhibited excellent cycling stability (%
capacity retention after  cycles). Further, a multi-level strat-
egy was proposed, which combines a core–shell structure, sur-
face AlOcoating, and bulk Ti doping to alleviate the structural
degradation and sluggish ion-diusion kinetics of a Ni-rich ox-
ide cathode (NaNi.Co.Mn.O).[ ] The introduction of Ti im-
proved the ion-diusion dynamics without reducing crystallinity.
In addition, the electrochemically active Ni ions were enriched in
the inner core to provide high capacity, while the Mn-rich shell
and thermally stable AlOcoating inhibit surface parasitic re-
actions and prevent the growth of microcracks to maintain the
structural integrity. Therefore, the modified cathode exhibited a
wide operating-temperature range ( to  °C) and excellent
sodium storage performance (. mA h gat  C; .%
capacity retention after  cycles at .C) (Figure c). There-
fore, the introduction of these oxide coatings physically protects
the active material from electrolyte attack, thereby maintaining
the structural integrity and electrochemical activity of the cath-
ode material.
It was reported that heat treatment can enhance the contact
between the coating and bulk material and also promote the
doping of some of the metal ions into the crystal lattice of the
cathode.[ ] Sucient evidence indicates that doping or substi-
tution of cations (Mg, Cu, Ti, Sn, and others) can eectively
maintain the stability of the layered framework, reduce the dif-
fusion barrier, and alleviate/suppress phase transitions.[,,]
Therefore, the integrated doping–coating strategy eectively en-
hances the electrochemical performance of the modified cath-
ode. It was reported that the increase in entropy mixing is a
driving force for the random substitution of Ni and dopants
(for example, Mg) in the lattice, enabling the synthesize of cath-
odes with a surface coating and bulk substitution via a one-
pot method.[ ] Accordingly, an O-Na[Ni.Mn.]Ocathode
coated with MgO and doped with Mg was demonstrated, in
which the coating layer minimizes parasitic reactions between
the active material and the electrolyte to prevent HF attacks,
while Mg+doping eectively reduces irreversible phase transi-
tions to improve structural stability (Figure d). A TiO-coated
O-NaMn.Fe. Ni. Ocathode was synthesized by dry mix-
ing followed by high-temperature calcination, and it was con-
firmed that some of the Ti+was doped into the lattice us-
ing cross-sectional EDS mapping and X-ray photoelectron spec-
troscopy depth profiling.[ ] The coating layer prevented surface-
side reactions to improve structural integrity, while Ti+doping
shortens the TM–O bond length and expands the interlayer dis-
tance to enhance the rate capability. A strategy combining bulk
Sn+substitution and surface nanolayer Sn/Na/O composite
Adv. Energy Mater. 2023, 2301975 © 2023 Wiley-VCH GmbH
2301975 (21 of 41)
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202301975 by Hanyang University Library, Wiley Online Library on [18/08/2023]. 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
www.advancedsciencenews.com www.advenergymat.de
Figure 10. a) HRTEM images of Al2O3-coated NaNi0.5Mn0.5 O2. Reproduced with permission.[243 ] Copyright 2021, Elsevier. b) Schematic diagram of
CEI thickness and Na-ion transport depending on the Al2O3coating. Reproduced with permission.[245 ] Copyright 2017, the Royal Society of Chemistry.
c) Cycling stability of the coated and compared samples at 0.5 C (1.5–4.1 V). Reproduced with permission.[237] Copyright 2023, Wiley-VCH GmbH.
d) Schematic diagram of the synthesis process of MgO-coated Na[Ni0.5Mn0.5 ]O2(top). TEM image and corresponding energy-dispersive spectroscopy
(EDS) spectra of a coated sample (bottom). Reproduced with permission.[236 ] Copyright 2018, the Royal Society of Chemistry. e) Cycling stability of
NaNi1/3Fe1/3Mn1/3 O2and Sn doped samples at 0.5 C. Reproduced with permission.[144 ] Copyright 2022, the American Chemical Society.
coating was shown to enhance the Na-ion storage performance
of O-NaNi/Fe/ Mn/Ocathodes.[ ] The lifetime of the Sn-
modified cathode was doubled (Figure e). This was attributed
to the surface coating reducing undesirable parasitic reactions,
Sn doping inhibiting adverse high-voltage phase transforma-
tion to improve structural reversibility, expanding the interplanar
spacing to enhance the ion diusion coecient, and increasing
the ionicity of the TM–O bond to increase the average operat-
ing voltage. In addition, the combined coating–doping strategy
not only extends the service life by improving structural integrity
and reducing internal reaction resistance, but also eectively im-
proves air stability and anion redox reversibility.[,] However,
it should be noted that this dual modification strategy requires
further optimization as it is still dicult to accurately control the
amount and location of the dopant.
In addition to improving chemical stability, the introduced pro-
tective coating eectively improves the environmental stability of
layered oxides by physically isolating the sensitive cathode from
moist air.[] The air stability of NaNi.Mn. Owas enhanced
using surface coatings of CuO[ ] or AlO.[  ] The chemical ag-
ing results indicated that this surface-modification strategy can
eectively alleviate the structural degradation (reduce the con-
tent of hydrated compounds), inhibit the formation of impurities
(NaHCOand NaCOby-products), and suppress the dissolu-
tion of TM ions. Manthiram et al. used inert ZrOto modify the
surface of NaNi.Mn.Co.Oto suppress its intercalation re-
action with HO/CO.[] Electrochemical tests showed that the
modified cathode delivered % of the initial capacity after days
of exposure to air, while the unmodified anode completely failed.
Although the aged cathode can be restored by high-temperature
annealing, this increases production costs.[ ] In summary, sur-
face engineering and element substitution strategies remain key
to improving environmental stability and achieving commercial
applications of layered-oxide cathode materials.
3.3.3. Polyanionic Composite Coatings
Polyanionic composites are considered suitable coating mate-
rials because of their potentially high ionic conductivity and
Adv. Energy Mater. 2023, 2301975 © 2023 Wiley-VCH GmbH
2301975 (22 of 41)
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202301975 by Hanyang University Library, Wiley Online Library on [18/08/2023]. 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
www.advancedsciencenews.com www.advenergymat.de
Figure 11. a) Schematic diagram of the synthesis process of an AlPO4-coated Na[Li0.05Mn0.50 Ni0.30Cu0.10 Mg0.05 ]O2cathode and the correspond-
ing TEM–EDS map of the AlPO4layer. Reproduced with permission.[252] Copyright 2019, Elsevier. b) Schematic diagram of the protective effect of
Na33xAlxPO4coating on active materials during repeated cycles and HRTEM images of the coated sample. Reproduced with permission.[ 240 ] Copyright
2023, the American Chemical Society. c) Structural model and corresponding TEM images of a Ti-rich spinel coating on a Mn-rich layered cathode. Repro-
duced with permission.[257 ] Copyright 2017, Springer Nature. d) Schematic diagram of an O3/O’3-P2 core–shell structured cathode and corresponding
surface and cross-sectional SEM images. Reproduced with permission.[258 ] Copyright 2020, American Chemical Society. e) Schematic diagram of the
synthesis process of P2-coated O3-structured cathode and a corresponding TEM image of the coating. f) Cycling stability of the pristine and coated
cathodes at 1 C. Reproduced with permission.[259 ] Copyright 2022, Elsevier.
stable framework structure.[ ] In addition, compared to com-
mon mechanical mixing methods for synthesizing oxide coat-
ings, polyanionic coatings are typically prepared by wet-chemical
methods, which facilitate the formation of uniform and well-
adhered layers.[ ] A thin AlPOion conductor was coated on
the surface of an O-Na[Li.Mn.Ni.Cu. Mg.]Ocathode
to enhance the ionic conductivity and alleviate the dissolution of
Mn (Figure 11a), thereby increasing the capacity retention from
% to % after  cycles at C.[] In addition, NaSiOand
NaTi(PO)ion conductors with D Na-ion diusion channels
Adv. Energy Mater. 2023, 2301975 © 2023 Wiley-VCH GmbH
2301975 (23 of 41)
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202301975 by Hanyang University Library, Wiley Online Library on [18/08/2023]. 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
www.advancedsciencenews.com www.advenergymat.de
were developed to protect the host material and reduce the migra-
tion energy barrier, thereby improving the rate capability.[, ]
A new type of ionic conductor NaZrSiPO coating was de-
posited on the NaNi/Mn/Fe/Omodel cathode by a sol-gel
process, followed by high-temperature calcination.[ ] The pro-
tective coating provided fast Na-ion diusion channels and re-
duced cracking and particle peeling. In addition, density func-
tional theory and neutron powder diraction analyses showed
that some of the Zr atoms were successfully doped into the TMO
layer, which expanded the Na-layer spacing, increased the self-
diusion coecient, and delayed the phase transition. Therefore,
the interface-optimized cathode exhibited improved rate capabil-
ity ( mA h gat C) and reduced electrochemical polariza-
tion (. V). Considering that one of the main factors lead-
ing to rapid capacity degradation is the residual Na-based im-
purities, such as NaOH and NaCOwith poor conductivity, re-
ducing or removing such surface compounds to improve struc-
tural stability is an eective modification strategy for layered ox-
ide cathodes.[, ] Therefore, simultaneously consuming sur-
face residues and applying a protective coating in situ is an ef-
fective method for overcoming the shortcomings related to Na
residues and interfacial instability.[] The in-situ formation of
a plastic-crystal NaxAlxPOcoating on the outer surface of an
O-NaNi.Fe. Mn.Ocathode was demonstrated using a wet-
chemical method, while a thin Na-deficient phase was formed
on the inner surface due to the extraction of Na from the crys-
tal lattice.[ ] This protective layer formed by consuming resid-
ual alkali compounds on the surface inhibits electrolyte corro-
sion and the accompanying parasitic reactions, isolates the sen-
sitive active materials from humid air, and provides a rapid dif-
fusion channel for Na-ions based on the paddle-wheel mech-
anism associated with coupled anion rotation and cation mi-
gration (Figure b).[] This strategy of coupling surface en-
gineering and Al doping can eectively alleviate the generation
of microcracks and particle fragmentation, resulting in an opti-
mized cathode with excellent rate performance (. mA h gat
C) and cycling stability (capacity retention increased from %
to % after  cycles at C). Recently, the surface residual
Na of O-NaNi.Cu.Mn. Ti.Owas transformed into a solid-
electrolyte NaMgPOcoating to improve the rate capability of
the modified cathode.[ ] The introduction of protective coat-
ings with fast ion-transport channels inhibits the dissolution of
the active material and improves the reversibility of the P–OP
phase transition within .–. V. Therefore, the resulting full-
cell showed a reversible capacity of  mA h gand a high
capacity retention of % after  cycles at . C. To date, this
strategy of removing surface residues while constructing a pro-
tective surface layer has not yet been fully developed, and further
optimization is required.
3.3.4. Heterostructure Coating
Although heterostructured cathode materials have been proven
to have excellent electrochemical performance, Mn segregation
to the particle surface is the main cause of capacity loss dur-
ing cycling.[ ] Contrastingly, coupling additional phases on the
surface of secondary particles enhances the direct interaction
between heterostructures.[ ] It has been shown that the intro-
duction of inactive protective coatings can eectively improve
the structural stability and environmental stability of layered
oxides, but at the expense of some capacity.[,] Therefore,
researchers have developed advanced core–shell cathode mate-
rials by coating fast ion/electron-conducting structural materi-
als (such as the P phase) on high-capacity structural materi-
als (for example, the O phase). The coating protects the sensi-
tive core material from the corrosive electrolyte and provides a
buer layer to reduce the eect of lattice changes in the cathode
during cycling. For example, a multiphase NaMn.Ti.Ni.O
cathode containing P and O’ layered core and a spinel-like
titanium(III)-oxide interface was fabricated using Ti-enrichment-
induced surface reconstruction (Figure c).[] The introduc-
tion of an atomically thick Ti-rich spinel-like interface pre-
vents the direct contact of the layered oxide with the moist
air/electrolyte and improves the electronic/ionic conductivity, so
the resulting cathode material exhibits excellent cycling stabil-
ity (% capacity retention after  cycles at C). To elimi-
nate surface NaCO, a thermal activation process was demon-
strated, which converts surface residual Na into an electrically
active P/O composite structure, thereby increasing the initial
charging capacity of NaxMnO(from  to  mA h g)de-
spite the dominant P phase leading to an abnormal ICE.[ ]
Furthermore, an O/O’-P core–shell structured cathode ma-
terial (Na.[(Ni.Co.Mn. ).(Ni. Mn.). ]O) was prepared
by a co-precipitation method, which benefited from the high-
capacity O-type core and stable P-type shell (Figure d).[]
The precursor with an Mn-rich surface is favorable for the for-
mation of the P phase, so the P-phase shell on the spher-
ical particles can alleviate the mechanical damage caused by
volume changes and provide fast Na-ion diusion paths. Al-
though introducing P coatings with Na vacancies can slightly
increase the discharge capacity, it can also lead to an abnor-
mal ICE. Hence, a -nm P-Na/MnOlayer was successfully
coated on an O-NaNi.Mn.Ocathode using a wet-chemical
process (Figure e).[] By carefully adjusting the thickness of
the P coating, the optimized cathode ( mol% coating) exhib-
ited a slightly higher ICE (%) and excellent cycle stability (%
capacity retention after  cycles at C) without sacrificing the
reversible capacity ( mA h g) (Figure f).
3.3.5. Other Coatings
In addition to the modified layers discussed above, researchers
have developed various coating materials to improve the struc-
tural stability of layered cathode materials. For example, organic
polymer coatings with mechanical flexibility and appropriate
electrical conductivity have been developed.[ ] For example, a
composite organic solvent “cocktail was used to form an artificial
CEI layer composed of metal–organic compounds and reduced
TM cations on the surface of NaNi/Fe/ Mn/Ocathodes to en-
hance the electrochemical stability of active materials in a propy-
lene carbonate (PC) based electrolyte, especially at high current
rates.[ ] A surface-modification strategy for a methacrylic-acid–
acrylonitrile copolymer was shown to improve the interfacial sta-
bility of P/O-Na.Li.Ni.Mn. O+𝛿cathodes under high
potentials.[ ] The introduced polymer with a strong electron-
donating group shares electrons with oxidized TM ions to
Adv. Energy Mater. 2023, 2301975 © 2023 Wiley-VCH GmbH
2301975 (24 of 41)
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202301975 by Hanyang University Library, Wiley Online Library on [18/08/2023]. 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
www.advancedsciencenews.com www.advenergymat.de
Figure 12. a) SEM image of O3–Na[Ni0.60Fe0.25Mn0.15 ]O2. Reproduced with permission.[92] Copyright 2020, Elsevier. b) SEM images of un-doped (left)
and Ti-doped (right) Na[Ni0.6Co0.2 Mn0.2 ]O2. Reproduced with permission.[180 ] Copyright 2019 Elsevier. c) SEM images of NaMn0.45Ni0.45Mg0.05 Ti0.05O2.
Reproduced with permission.[91] Copyright2022 Elsevier. d) Design scheme and corresponding cross-sectional TEM image of radially aligned hierarchical
columnar particles. Reproduced with permission.[265 ] Copyright 2015, Springer Nature. e) SEM and TEM images of Na[Li0.05Mn0.50 Ni0.30Cu0.10 Mg0.05 ]O2
cathode material. Reproduced with permission.[268 ] Copyright 2017, Wiley-VCH GmbH. f) Schematic diagram of the synthesis process of core–shell
structured cathode materials. Reproduced with permission.[154 ] Copyright 2020, Wiley-VCH GmbH. g) Schematic diagram of the synthesis process
of hollow microbar-like NaNi0.5Mn0.5 O2cathode and corresponding typical TEM image. Reproduced with permission.[269 ] Copyright 2020, Elsevier.
h) SEM images of cubic-like NaMn0.48Fe0.32Cu0.2 O2. Reproduced with permission.[198 ] Copyright 2023, Elsevier.
anchor them and inhibit the dissolution of the active material.
In addition, the slightly higher valence of O caused by a charge-
balancing mechanism leads to an increase in the lattice-O activity
and Na-ion diusion rate. Furthermore, an AlF-coated nanorod
O-Na[Ni.Co.Mn.]Omaterial was developed to limit the
structural degradation associated with the formation of a thick
rock-salt layer and inhibit the mechanical failure caused by the
growth of intra-particle cracks.[ ] Therefore, the pouch-type full
cell assembled with this optimized cathode and HC anode exhib-
ited enhanced cycling stability, with a high capacity retention of
% after  cycles at . C.
3.4. Optimal Synthesis Process: Co-Precipitation versus Other
Methods
Coprecipitation is a common method for preparing cathode ma-
terials with controllable morphology and high tap density for
both LIBs and SIBs.[, ] The method usually includes the syn-
thesis of precursor particles (hydroxide, oxalate, or carbonate)
by a coprecipitation reaction. The precursors are then mixed
with a Na source and processed by high-temperature calcina-
tion. Spherical precursors with a diameter of several microns are
usually obtained by the coprecipitation method and their mor-
phology is inherited by the final product after calcination.[, ]
In addition, because the calcination temperature of the sodi-
ated oxide cathode is usually higher than that of the Li analogs,
grain-boundary fusion can occur. Therefore, polyhedral spheri-
cal cathode materials with dense surfaces have been reported
(Figure 12a).[] Conversely, the introduction of dopants can af-
fect the rate of crystal growth and ultimately alter the morphol-
ogy of the product.[ ] For instance, it was demonstrated that
the introduction of Ti as a dopant can induce the growth of pri-
mary particles during heat treatment, which resulted in the syn-
thesis of densely packed secondary particles with a lower void
volume and enhanced mechanical strength (Figure b).[] Be-
cause Mg/Ti co-doping accelerates the growth of the O struc-
ture, the final product exhibited spherical and plate-like mor-
phologies (Figure c).[] To further improve the Na-ion storage
performance, researchers have prepared cathode materials with
unique morphologies by adjusting the synthesis conditions. In
, a new concept of assembling spherical particles using a
radially aligned hierarchical columnar structure was presented,
and the product was characterized by a concentration gradient
of TM ions (Figure d).[] The closely arranged secondary par-
ticles minimize the contact between sensitive Ni and corrosive
Adv. Energy Mater. 2023, 2301975 © 2023 Wiley-VCH GmbH
2301975 (25 of 41)
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202301975 by Hanyang University Library, Wiley Online Library on [18/08/2023]. 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
www.advancedsciencenews.com www.advenergymat.de
Figure 13. SEM images of a1) pristine and a2) Sb-doped NaNi0.5Mn0.5 O2. Reproduced with permission.[270 ] Copyright 2022, American Chemical Soci-
ety. b) SEM images of NaNi0.5Mn0.3Co0.2 O2with hexagonal plate-like morphology. Reproduced with permission.[273] Copyright 2020, IOP Publishing.
c) SEM images of Na0.5Mn0.65 Ni0.15Al0.1 Mg0.05 Co0.05O2cathode material with nano-flower structure. Reproduced with permission.[274] Copyright 2020,
Springer Nature. d) SEM and TEM images of Na[Li0.05Ni0.3 Mn0.5Cu0.1 Mg0.05 ]O2with exposed (010) active facets. Reproduced with permission.[278 ]
Copyright 2018, Wiley-VCH GmbH. e) SEM images of hollow spherical NaNi1/3Mn1/3 Fe1/3 O2structures. Reproduced with permission.[189 ] Copyright
2022, Elsevier. (f) SEM images of hollow fiber Na3Ni2SbO6cathode material. Reproduced with permission.[279 ] Copyright 2022, Wiley-VCH GmbH.
electrolyte, thus achieving extended service life (% capacity
retention after  cycles) and improved thermal stability (in-
creasing the exothermic peak by  °C). To highlight the ad-
vantages of this structure, the mechanical and electrochemical
characteristics of the cathode composed of spoke-like nanorods
were compared with those a constant-concentration sample.[ ]
The results of micro-compression tests and high-resolution TEM
reveal that the compact structure can enhance the mechanical
strength by minimizing the porosity, and improve the ionic dif-
fusion dynamics through the directional transmission of elec-
trons/ions. In addition, radially distributed primary particles can
be synthesized in other constant-concentration cathode materi-
als, which indicates that the coprecipitation conditions have a
great impact on the particle morphology.[] By optimizing the
synthesis process (pH, temperature, and feeding rate), a sub-
micron spherical precursor was formed by the aggregation of
nanoplate primary particles, which exhibited high industrial fea-
sibility (Figure e).[] Due to the short Na-ion diusion dis-
tance, the obtained cathode material showed excellent rate prop-
erties. In addition, O-type NaNi.Mn.Ocathodes with a core–
shell structure were developed by adjusting the chemical compo-
sition, in which the Ni-rich core provides high capacity, and the
Mn-rich shell improves the structural stability (Figure f).[]
However, the core–shell structured cathode exhibited unsatisfac-
tory cycling stability, which may be related to the formation of
irregular flaky particles. In addition to conventional spherical
precursors, coprecipitation can be used to synthesize cathodes
with unique morphologies by adjusting the solvents and chelat-
ing agents. For instance, hollow O-NaNi.Mn.Omicrobars
prepared using ethanol-mediated coprecipitation were composed
of oriented stacks of nanoplates with exposed () crystal facets
(Figure g).[] The resulting cathode material exhibited excel-
lent rate capability and an extended service life owing to the en-
hanced exposure of electrochemically active planes, which short-
ens the diusion distance and reduces volume changes during
the sodiation/desodiation process. A dispersed microscale O-
NaMn. Fe. Ocathode material was synthesized using a cubic
Mn[Fe(CN)]·HO precursor (Figure h).[] Although us-
ing PBAs as templates can eectively control the morphology of
the final product, this method can only be used to synthesize cath-
ode materials with specific chemical components.
Solid-state fabrication is one of the traditional methods for
synthesizing cathode materials, which has the advantages of
simple processing steps and accurate product stoichiometry.
Nanoscale O-NaNi.Mn.Owith controllable grain orienta-
tion was achieved by Sb substitution using a solid-state method
(Figure 13a).[] The introduced dopant can refine the grains by
modifying the surface energy of layered cathodes,[ ] broaden
Adv. Energy Mater. 2023, 2301975 © 2023 Wiley-VCH GmbH
2301975 (26 of 41)
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202301975 by Hanyang University Library, Wiley Online Library on [18/08/2023]. 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
www.advancedsciencenews.com www.advenergymat.de
the Na-ion transport channels, and reduce the ion-diusion en-
ergy barrier. Therefore, the optimized cathode delivered an out-
standing rate capability of . mA h gat  C. The solid-state
synthesis process is relatively simple, but the subsequent long
high-temperature sintering process ( to  °C) consumes
a lot of energy, resulting in low processing eciency.[] Al-
though the sol-gel method can reduce the sintering temperature
( °C), the final product is usually composed of agglomerated
nano/microplates, which limits the transfer of ions/electrons
during the charge/discharge process.[ ] Similarly, a solution
combustion synthesis technique was proposed to prepare chem-
ically uniform cathode materials.[ ] Specifically, nitrate and
glycine were used as the oxidant and fuel, respectively, to ob-
tain micron-sized secondary particles composed of hexagonal
plate-like primary nanoscale particles by auto-ignition at  °C
(Figure b).[] However, the electrochemical performance of
the resulting cathode was not satisfactory, which may be due to
its loose stacking. In addition, the consistency of the morphology
of products prepared by sol-gel and solution-combustion meth-
ods is worse than that of products prepared by coprecipitation,
which is not conducive to large-scale production.
Hydrothermal methods are also eective for synthesizing ac-
tive materials with controllable morphology. The synthesis of a
nano-flower quinary cathode material was achieved using a urea-
assisted hydrothermal method, combined with subsequent high-
temperature calcination (Figure c).[] The resulting cath-
ode material provided surprisingly long cycling stability, with
a high capacity retention of .% after  cycles at C.
By controlling the synthesis conditions, a series of thin sheet-
like structures and nano/micro-spherical cathode materials have
been developed, which demonstrated enhanced electrochemical
performance.[– ] In addition, appropriate crystal facet design
to expose more active sites is an eective strategy to improve
the rate capability of layered cathodes. A five-component cathode
material (O-Na[Li.Ni. Mn.Cu. Mg.]O)withanexposed
() crystal facet was synthesized by thermal polymerization,
which comprised multilayers of oriented stacked nanosheets
(Figure d).[] By increasing the exposed electrochemically ac-
tive plane to shorten the diusion distance and maintain high
electrochemical reversibility through multi-element substitution,
the resulting cathode material exhibited outstanding rate perfor-
mance (. mA h gat  C) and long-term cycling stability
(.% of the initial capacity after  cycles). Although the prod-
uct synthesized by a hydrothermal method showed unique mor-
phology and enabled subsequent calcination at moderate temper-
atures, the feasibility of upscaling this method needs to be con-
firmed.
Micro-spherical cathode particles assembled by primary par-
ticles can also be obtained by spray drying. For instance, TM-
oxides were used as raw materials to prepare spherical Ni/Mn/Fe-
based cathode precursors by a combined ball milling–spray dry-
ing process (Figure e).[,] The obtained porous hollow prod-
ucts are beneficial for electrolyte permeation and exhibit im-
proved ion-diusion kinetics. Although rapid spray-drying meth-
ods have commercial potential, they still require further op-
timization to obtain uniform cathode materials. In addition,
due to the loose characteristics of the precursors, it is neces-
sary to avoid particle cracking during high-temperature calcina-
tion to further increase tap density and structural integrity.[]
Electrospinning to synthesize one-dimensional nanofibers with
uniformly adjustable diameters and interconnected framework
structures is an eective strategy for achieving a high rate perfor-
mance and low internal resistance of cathode materials.[, ]
Hollow-fiber NaNiSbOcathode nanoparticles were fabricated
using such a method (Figure f).[] Compared with equivalent
materials with an irregular morphology, the fiber-based cathode
showed improved rate performance due to exposed active sites
and reduced volume changes. However, the subsequent high-
temperature calcination process in air may lead to the growth
of primary nanoparticles and interrupt the ion/electron trans-
port paths. Thus, the practicality of electrospinning technology
for producing oxide cathode materials needs further analysis.
3.5. Single-Crystal Particles
It should be noted that many grain boundaries exist in the sec-
ondary particles composed of primary particles, which induce
the formation of microcracks that result in capacity decay and
thermal instability during the charging/discharging process.[]
In LIBs, the single-crystal cathode without inter-particle bound-
aries has attracted the attention of academia and industry, be-
cause it can avoid the formation of cracks to obtain extended
service life.[ ] Based on this, a solvothermal method was used
to prepare a single-crystal O-NaFeOcathode material with
a hexagonal morphology.[] Furthermore, micron-sized trun-
cated octahedra or platelet-like single-crystal particles were ob-
tained by molten salt-assisted synthesis.[, ] First, the synthe-
sis conditions were optimized to produce Na(Ni.Fe. Mn.)O
single crystals with enhanced cycling stability. These modifica-
tions included using metal oxides instead of coprecipitation par-
ticles as the starting materials, % molten-salt (NaOH) con-
tent, a temperature of  °C, and rinsing the particles in wa-
ter (Figure 14a).[] To avoid parasitic reactions between the
alumina crucible and NaOH flux, a two-step synthesis method
was further proposed, which uses NaCl as the molten salt and
metal oxides as raw materials to prepare single-crystal trun-
cated octahedral NaNi.Mn.Oparticles with high tap density
(. g cm).[ ] The resulting single-crystal cathode with high
crystallinity and consistent lattice orientation eectively mini-
mized surface parasitic reactions, reduced the generation of in-
tergranular cracks, and exhibited a longer lifetime than polycrys-
talline particles (Figure b). Although single-crystal particles ex-
hibit improved cycling stability, there are still some issues that
need to be addressed. For example, synthesizing single-crystal
cathodes by molten-salt methods requires an additional water-
washing step, which adds complexity to the production process
and may induce structural transformation of the active mate-
rial. Therefore, it is necessary to develop alternative methods
to simplify the synthesis process.[ ] Due to the lengthening of
Na-ion diusion pathways and reduced electrolyte accessibility,
large single-crystal particles typically exhibit slightly reduced re-
versible capacity and sluggish reaction kinetics (rate capability),
which require further optimization. In addition, since selected
area electron diraction analysis is confined to a narrow area, it is
not suitable to prove that micron/submicron particles are single
crystals.[ ] Therefore, it is recommended that electron backscat-
ter diraction analysis is conducted as this method can detect
Adv. Energy Mater. 2023, 2301975 © 2023 Wiley-VCH GmbH
2301975 (27 of 41)
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202301975 by Hanyang University Library, Wiley Online Library on [18/08/2023]. 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
www.advancedsciencenews.com www.advenergymat.de
Figure 14. a) Cycling performance of Na(Ni0.3Fe0.4 Mn0.3 )O2single crystals synthesized by a molten-salt method with different raw materials and rinsing
solvents. Reproduced with permission.[284 ] Copyright 2022, Wiley-VCH GmbH. b) Cycling stability and corresponding SEM images of single crystal (MS-
SNMO) and polycrystalline (PC-SNMO) NaNi0.5Mn0.5 O2. Reproduced with permission.[285 ] Copyright 2022, American Chemical Society.
grain boundaries and provide direct evidence for the synthesis of
single crystals. Although the research on single-crystal cathode
materials for SIBs is still in its early stages and the cycling stabil-
ity of these materials is slightly inferior to some modified poly-
crystalline cathodes,[,, ] it is proposed that with further re-
search eorts, the electrochemical performance of single-crystal
cathodes will be significantly improved soon.
4. Anodes and Electrolytes for Use with Practical
Cathodes
4.1. Hard-Carbon Anodes
The anode materials for SIBs mainly include C-based materials,
alloying materials (P, Sn, Sb, and others), and metal oxide/sulfide
materials.[] Although anodes that undergo alloying and conver-
sion reactions can deliver high reversible capacity, their large vol-
ume expansion during sodiation leads to the crushing and struc-
tural collapse of the active material, which greatly limits the cy-
cling life of SIBs.[ ] Furthermore, the high reversible capacity
of an anode material does not directly correspond to high-energy
batteries, which is also closely related to their average charg-
ing voltage.[ ] Titanium-based oxide/phosphate anode materi-
als based on intercalation reactions have attracted widespread
attention because of their excellent cycling stability and rate ca-
pacity, and small volume changes. However, their high operat-
ing voltage and low specific capacity limit the energy density of
the full-cell system. Therefore, insertion-based carbonaceous ma-
terials are considered the most promising anode materials for
commercial SIBs because of their diverse structures, high nat-
ural abundance, and low cost. Carbonaceous materials include
graphite, soft carbon (SC), HC, CNTs, and graphene. Graphite is
Adv. Energy Mater. 2023, 2301975 © 2023 Wiley-VCH GmbH
2301975 (28 of 41)
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202301975 by Hanyang University Library, Wiley Online Library on [18/08/2023]. 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
www.advancedsciencenews.com www.advenergymat.de
Figure 15. a) SEM images of a cork-derived HC anode. Reproduced with permission.[297 ] Copyright 2019, Wiley-VCH GmbH. b) Schematic diagram
of a phosphorus-functionalized HC anode with enhanced electrochemical performance. Reproduced with permission.[300 ] Copyright 2018, Wiley-VCH
GmbH. c) Rate performances of control samples and a carboxyl-doped HC anode (C1600-M). Reproduced with permission.[301 ] Copyright 2021, Wiley-
VCH GmbH. d) ICE as a function of various C characteristics (CO2specific surface area, COxgroups, and active surface area). Reproduced with
permission.[302 ] Copyright 2021, the Royal Society of Chemistry. e) Specific capacity and ICE changes of HC anodes under the catalytic effects of various
metal ions. Reproduced with permission.[303 ] Copyright 2023, Wiley-VCH GmbH. f) Galvanostatic charge/discharge curves (1st, 2nd, 10th, and 100th
cycles) and corresponding structural diagrams of various HC paper samples. Reproduced with permission.[304 ] Copyright 2021, American Chemical
Society.
currently used as the anode material for commercial LIBs. Al-
though graphite with solvent co-intercalation mechanism pro-
vides excellent cycling stability and rate capability in SIBs sys-
tems, its low specific capacity ( mA h g) and high operat-
ing voltage (. V versus Na+/Na) are not conducive to achieving
a high energy density in the full cell, which greatly limits its appli-
cation potential.[ ] Compared to the high price and low packing
density of CNTs and graphene, as well as the low ICE and high
voltage plateau of SC, HC with high reversible capacity (–
 mA h g) and low discharge voltage (. V versus Na+/Na)
is considered the most promising anode material for commer-
cial SIBs, which can be demonstrated through the research path
of SIB startups.[ ]
4.1.1. Na Storage Performance
The electrochemical behavior of Na ions in HC anodes involves
adsorption on surfaces and defects, intercalation reactions, and
filling of micropores.[ ] At present, the main controversy re-
garding the Na storage mechanism of HC anodes is the con-
tribution of various reactions to the slope regions (above . V)
and plateau region (below . V). To achieve commercial appli-
cations of SIBs, various challenges related to HC anodes still
need to be addressed, including their poor rate capability and
unsatisfactory cycling stability caused by slow ionic diusion
in the low voltage plateau region, as well as diculties in con-
structing practical full cells due to the low ICE. Shortening
the Na-ion transfer distance by morphological engineering can
eectively improve the poor rate capability of HC materials. For
example, porous HC fibers synthesized by electrospinning and
subsequent high-temperature carbonization processes can eec-
tively enhance the electrolyte accessibility and ionic diusion
eciency.[,] In addition, C-microspheres synthesized by hy-
drothermal methods or hollow C-spheres prepared by templat-
ing methods (with templates of SiO, TiO, and others) not only
shorten the ionic diusion paths, but also avoid the formation of
uneven and thick solid electrolyte interphase (SEI) layers, thereby
achieving high reversible specific capacity and improved Na stor-
age performance of the obtained HC anode.[, ] HC mate-
rials with nanoporous networks and wide interlayer spacings
were synthesized using a MgO-templating method, which ex-
hibited low operating voltage (. V), high reversible capacity
( mA h g), and appropriate ICE (%).[ ] HC materials
synthesized by carbonization of inexpensive and easily available
biomass-derived materials, such as discarded fruit shells, cotton,
poplar, and straw have attracted the attention of researchers ow-
ing to their low cost and unique inheritable structure. HC anodes
with a hierarchical porous structure were fabricated from waste
cork (Figure 15a) and exhibited a reversible capacity of  mA
hg
and an ultra-long lifespan of over  cycles (% ca-
pacity retention) in a Na half-cell.[] Although morphological
Adv. Energy Mater. 2023, 2301975 © 2023 Wiley-VCH GmbH
2301975 (29 of 41)
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202301975 by Hanyang University Library, Wiley Online Library on [18/08/2023]. 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
www.advancedsciencenews.com www.advenergymat.de
engineering can eectively improve the reversible capacity and
rate capability of HC materials, an increase in specific sur-
face area may lead to poor cycling stability and high initial
irreversibility.[] In addition, the nanoscale design of anode
materials may lead to processing diculties (including mate-
rial synthesis and electrode manufacturing), which is not favor-
able for industrial production. Moreover, the C-conversion ef-
ficiency of HC materials formed by the carbonization of poly-
mer/biomass materials should also be considered to evaluate
their feasibility.[,] Therefore, more eorts are required to op-
timize the synthesis process to balance the manufacturability,
morphology, and electrochemical performance of HC materials,
which is the key to promoting the commercialization of SIBs.
Dopingwithheteroatoms(suchasN,S,O,P,B,andF)is
also an eective strategy for improving the Na-ion storage per-
formance of HC materials by changing the electronic state of C
to increase the electronic conductivity while also increasing the
number of defect/surface functional groups to increase the re-
versible capacity.[] For example, it was demonstrated that phos-
phorus functionalization of HC can broaden the interlayer spac-
ing and increase the Na-adsorption energy, resulting in a high
initial reversible capacity (. mA h g) and excellent rate per-
formance (. mA h gat A g) of the resulting anode
(Figure b).[] Furthermore, a multistage porous HC anode
with a high concentration of pyridine N was formed by polymer-
ization and pyrolysis using tri-sodium citrate (C-source and tem-
plate) and hexamethylenetetramine (N source and pore-forming
agent) as raw materials.[ ] The introduction of C–Nand
C–Cradicals provide additional active sites for the adsorption of
Na ions under low voltage, so the obtained anode shows high re-
versible capacity ( mA h g) and rate capability ( mA h g
at A g). A novel pre-oxidation treatment was proposed to form
a more disordered structure and introduce more O-functional
groups, thereby improving the reversible capacity (from  to
. mA h g), C-yield (from % to %), and ICE (from
.% to .%) of carbonized and pre-oxidized pitch.[ ] Simi-
larly, carboxyl groups were introduced into carbonized anthracite
by a mechanochemical post-treatment process to expand the in-
terlayer distance and enhance the porosity and disorder/defect
structure.[ ] As shown in Figure c, this modified HC anode
with a precise content of  at% carboxyl groups (C-M) ex-
hibited high reversible capacity ( mA h g) and excellent rate
performance ( mA h gat A g). It should be noted that
compared to traditional HC, modified HC materials with more
defects induced by heteroatom doping exhibit increased capac-
ity in the sloped region of the current–voltage curve, which may
lead to an increase in the average charging voltage and oset the
benefits of increased capacity.[,] In addition, introducing de-
fects and porous structures into HC materials can increase the
reversible capacity, but it may also lead to an increase in initial ir-
reversible capacity, making it dicult to fabricate practical Na-ion
full cells.[, ]
4.1.2. Increasing the ICE
The practical application potential of SIBs is limited by the low
ICE of the anode material caused by electrolyte decomposition
and side reactions, as all Na-ions are originally stored only in
the cathode, not in the HC anode. After years of development,
significant progress has been made in the electrochemical per-
formance of HC anodes, but the ICE of most reported HC ma-
terials is only about %, which does not yet meet the require-
ments for practical applications.[, ] Although pre-sodiation
treatment (by electrochemical methods, direct contact, or sacri-
ficial agents) can improve the performance of Na full cells, it in-
creases the number of processing steps and production costs.[]
Therefore, it is necessary to further optimize the HC anode to
meet the technical requirements for the commercial application
of SIBs. It should be noted that the service life and ICE of HC
materials are also aected by the electrolytes, binders, and con-
ductive agents, which require evaluation of the samples under
the same conditions to enable objective comparisons. For exam-
ple, selecting water-soluble binders (sodium carboxymethyl cel-
lulose, sodium polyacrylate, sodium alginate, and others) or re-
ducing the amount of super C with a large specific surface area
can improve the ICE of HC electrodes.[, ]
A systematic investigation of the relationship between the
defect concentration and ICE in HC materials demonstrated
that Na ions trapped by defect/surface functional groups gen-
erate electrostatic repulsion against other ions, thereby reduc-
ing the low-voltage intercalation capacity and increasing initial
irreversibility.[] Therefore, they proposed reducing the pyrol-
ysis rate to provide sucient time for gas removal and C-atom
reorganization, thereby increasing the ICE from .% to .%.
In addition, the pyrolysis temperature aects the defects, poros-
ity, and graphitization degree of HC materials.[] It is gener-
ally believed that a high carbonization temperature will lead to
the closure/collapse of pores and the reduction in the number of
defects, thereby increasing the ICE and low-potential plateau ca-
pacity of the obtained HC anode.[ ] However, HC synthesized
at high pyrolysis temperatures exhibits a high degree of graphi-
tization and narrow interlayer spacing, which degrades the Na
storage performance. By investigating the electrochemical perfor-
mance of HC anodes fabricated by direct pyrolysis of pitch and
phenolic resin at temperatures ranging from  to  °C,
the optimal carbonization temperature was determined.[ ] The
sample synthesized at  °C exhibited the highest ICE (%)
and improved cycling stability (% capacity retention after 
cycles at .C), owing to a balance between electronic conduc-
tivity (graphitization degree) and ionic conductivity (interlayer
spacing). In addition, research on the carbonization synthesis
of HC microspheres using phenolic resin precursors showed
that increasing the pyrolysis temperature can eectively enhance
the closed porosity and tap density.[] Therefore, a rolled HC
anode with low electrode porosity (%) and high tap density
(. g cmat  °C) exhibited an impressively high ICE of
% (Figure d). Furthermore, it was shown that graphite crys-
tal templates can induce the epitaxial growth of cotton precur-
sors under intimate contact, thereby forming an intergrowth HC
anode composed of graphite-like crystals, graphite phases, and
pseudo-graphite domains.[ ] The HC with graphite-like crystals
and wide interlayer spacing (. Å) exhibited an ultra-high ICE
(%), suitable reversible capacity ( mA h g), and a long
low-voltage plateau.
In addition to the carbonization temperature, the graphiti-
zation and defect degree of HC materials are influenced by
other conditions, such as the addition of transition metals with
Adv. Energy Mater. 2023, 2301975 © 2023 Wiley-VCH GmbH
2301975 (30 of 41)
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202301975 by Hanyang University Library, Wiley Online Library on [18/08/2023]. 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
www.advancedsciencenews.com www.advenergymat.de
catalytic graphitization ability. The graphitization degree of HC
anodes was carefully controlled by introducing appropriate con-
centrations of Mn ions into the precursor of paper towels.[ ]
Mn ion-assisted catalytic graphitization eliminates O-defects and
maintains eective ionic transport channels, thereby increasing
the ICE of HC samples from . to .% (Figure e). Simi-
larly, atomically dispersed Zn was added into HC microspheres
to catalyze the decomposition of the electrolyte and form a stable
inorganic-rich SEI layer.[] The optimized HC anode exhibited
a sucient ICE (%) and enhanced low-temperature capacity
( mA h gat  °C).
Considering the influence of the electrode interface behav-
ior on the formation and properties of SEI layer, surface-coating
technology is considered an eective strategy to enhance the ICE.
For example, an AlOlayer (artificial SEI) was evenly coated on
HC spheres by a simple liquid-phase method, which reduces the
surface active sites and inhibits parasitic reactions to increase the
ICE (from . to .%).[] Similarly, a vapor C-coating strat-
egy was demonstrated via the pyrolysis of polypropylene to en-
hance the ICE (%) of C-spheres, as the coating reduces surface
defects and fills the pores.[ ] HC anodes coated with SC were
fabricated by carbonizing paper towels soaked in coal-tar pitch
at  °C.[ ] The high H/C ratio of pitch reduces O-defects in
HC during carbonization and the long-range-ordered SC coating
avoids the formation of a thick SEI layer and subsequent con-
sumption of active Na-ions, resulting in the modified HC elec-
trode exhibiting a high ICE of .% (Figure f). However, most
surface-engineering technologies are still in the laboratory stage
and not yet suitable for large-scale application. Furthermore, they
have limitations related to the complex production steps and high
costs, requiring careful evaluation of their applicability.
4.2. Electrolyte Engineering
Electrolytes are an important component of electrochemical
energy-storage systems and largely determine the electrochem-
ical performance and safety of the batteries. An ideal electrolyte
should include the following characteristics: wide operating volt-
age window (large energy gap between the highest occupied
molecular orbitals (HOMO) and lowest unoccupied molecular or-
bitals (LUMO)), good compatibility with cathode and anode mate-
rials, high ionic conductivity, low viscosity, chemical/thermal sta-
bility, and low cost.[] Traditional non-aqueous electrolytes in-
clude sodium salts, organic solvents, and additives, as discussed
below.
Sodium salts should have high solubility and excel-
lent ionic conductivity, among which sodium perchlorate
(NaClO), sodium hexafluorophosphate (NaPF), and sodium
bis(trifluoromethane)sulfonimide (NaTFSI) have been widely
investigated. In PC-based electrolytes, NaPFexhibits the highest
conductivity (. mS cm), followed by NaClO(. mS cm)
and NaTFSI (. mS cm)(Figure 16a).[ ] It was demonstrated
that the thermal stability of electrolyte salts follows the order:
NaClO>NaTFSI >NaPF>Na bis(fluorosulfonyl)imide
(NaFSI).[] Due to its fast ion-transport rate, high thermal
stability, good compatibility, low water sensitivity, and low cost,
NaClOis widely used in prototype SIBs.[, ] However, its
toxicity and explosiveness in a dry state limit its large-scale
production, making it more popular in research than industry.
NaPFwith high ionic conductivity is a suitable sodium salt,
but it still faces challenges such as high toxicity, high price,
and low decomposition temperature. In addition, the solubility
of NaPFin a single solvent is limited, requiring the use of
mixed solvents.[ ] NaTFSI and NaFSI are also considered
promising sodium salts owing to their high ionic conductivity,
and non-toxicity, but their application is hindered by their severe
corrosion reaction with the Al current collector.[] Therefore,
NaPFprovides a suitable compromise in terms of electrochem-
ical stability, ionic conductivity, and safety, so it is considered the
best commercial sodium-salt candidate.[ ]
As a medium for Na-ion transport, organic solvents need to
meet the requirements of high dielectric constant (ɛ; conducive
to the dissociation of sodium salt), low viscosity (rapid ionic diu-
sion), high electrochemical stability (wide operating voltage win-
dow), high security (high boiling point and flash point), low tox-
icity, and low cost.[] Common organic solvents include car-
bonate (for example, PC, ethylene carbonate (EC), dimethyl car-
bonate (DMC), methyl ethyl carbonated (EMC), diethyl carbon-
ate (DEC)) and ether-based solvents (such as dimethoxyethane
(DME), diglyme, triglyme). Ether-based electrolytes can form
functional SEI layers on the surface of active materials during the
electrochemical process, so they show satisfactory electrochem-
ical performance in HC half-cells and Na metal batteries.[]
However, their low thermal stability (DME <DMC <DEC <PC <
EC) and incompatibility with high-voltage cathode materials limit
their application in Na full-cells.[] Organic electrolytes in SIBs
typically contain EC (ɛ=. at  °C) and PC (ɛ=. at  °C)
with high dielectric constants, although their viscosity is rela-
tively high (Figure b).[] However, EC exists in solid form at
room temperature (melting point . °C), so it is usually neces-
sary to mix one or more co-solvents with low viscosity (including
DMC, EMC, DEC) to optimize the electrochemical performance
of the resulting SIB. An EC:PC mixture with high thermal sta-
bility and a wide electrochemical stability window was proposed
as an optimal electrolyte formulation, but electrolytes based on
EC:DMC and EC:EMC have kinetic advantages.[ ] For example,
HC anodes exhibited high reversible capacity, low initial capacity
loss, and excellent cycling stability in EC:PC-based electrolytes
due to the formation of a thin and uniform SEI layer.[,]
In addition, high-voltage SIBs assembled with ternary solvent-
based electrolytes with high ionic conductivity and low viscos-
ity, such as EC/PC/DMC (././.), EC/PC/DEC (//),
and EC/DMC/DEC (//), exhibited excellent electrochemical
performance.[, ] However, there is currently no consensus on
the optimal solvent composition. Considering the influence of
the solvation structures of dierent electrolyte components on
battery performance, it is necessary to further optimize the sol-
vent design.[ ]
Because organic electrolytes decompose and corrode the elec-
trode during the electrochemical process, film-forming additives
such as fluoroethylene carbonate (FEC), vinylene carbonate (VC),
ethylene sulfite (ES), and difluoroethylene carbonate (DFEC) are
commonly added to extend the service life of the entire battery
to thousands of cycles.[ ] To date, FEC remains the most ef-
fective additive for SIBs. Adding a small amount of FEC (.%)
to NaPF-based electrolytes can significantly improve the re-
versible capacity, ICE, and cycling stability of HC anode materials
Adv. Energy Mater. 2023, 2301975 © 2023 Wiley-VCH GmbH
2301975 (31 of 41)
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202301975 by Hanyang University Library, Wiley Online Library on [18/08/2023]. 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
www.advancedsciencenews.com www.advenergymat.de
Figure 16. a) Conductivity and viscosity values of PC-based electrolyte with 1 мNa salts. Reproduced with permission.[315 ] Copyright 2012, the Royal
Society of Chemistry. b) Permittivity and viscosity of various organic solvents in SIBs. Reproduced with permission.[314] Copyright 2022, Wiley-VCH
GmbH. c) Cycling stability of HC electrodes in electrolytes with different compositions. Reproduced with permission.[322 ] Copyright 2016, Wiley-VCH
GmbH. d) Schematic diagram of the formation of CEI on the cathode after cycling in electrolytes with and without FEC additives. Reproduced with
permission.[323 ] Copyright 2021, Wiley-VCH GmbH. e) Effect of EFPN additive content on the flammability and ionic conductivity of NaPF6/EC–DEC
based electrolytes. f) Flammability test in a blank electrolyte (left) and electrolyte with 5% EFPN additive (right). Reproduced with permission.[ 324]
Copyright 2015, the Royal Society of Chemistry.
(Figure c), as it is preferentially reduced to form a functional
SEI layer containing NaF and Na polycarbonate.[] In addition,
electrolytes containing FEC additives induce the formation of a
F-rich CEI layer on the surface of the cathode material at high
voltage (. V) to promote Na-ion diusion (Figure d).[]
Although this uniform passivation layer slightly increases the
charge transfer resistance of the cathode, it eectively avoids
electrolyte consumption caused by the continuous decompo-
sition of oxidation products under high voltage, making elec-
trolytes containing FEC additives popular in SIBs systems.[, ]
A dual-additive electrolyte containing FEC and succinic anhy-
dride (SA) was used to improve the electrochemical performance
of Na.Li.Ni.Mn. Cu.Oand reduce the generation of
COduring electrolyte decomposition.[ ] The introduction of
SA induced the formation of a uniform and stable CEI layer,
which contains more O-rich organic compounds and less high-
resistance NaF compared to electrolytes containing only FEC ad-
ditives. In addition, compared with LIB systems, the Na-based
SEI layer composed of non-cross-linked or non-polymerized or-
ganic species is more easily dissolved, leading to an increase in
the self-discharge rate; thus, only a few LIB electrolyte additives
can be applied to SIBs.[ ] For example, the popular LIB elec-
trolyte additive VC is not highly eective in SIB systems.[]
To achieve high energy-density SIBs, it is necessary to fur-
ther develop stable high-voltage electrolytes, which can be
achieved by introducing new film-forming additives,[ ] flu-
orinated electrolytes,[ ] or high-concentration electrolytes.[]
More importantly, because of the inherent flammability of or-
ganic solvents, improving the thermal stability of electrolytes is
particularly important for improving the safety of SIBs, which
can be optimized by applying the following strategies: ) employ-
ing nonflammable solvents or adding flame-retardant additives;
) adopting high concentration/localized high-concentration
electrolytes; and ) developing suitable ionic liquids or solid
electrolytes.[ ] Commonly used phosphorus/fluoride-retardant
additives terminate combustion reactions by physically iso-
lating O via a flame-retardant gas vapor or chemically cap-
turing free hydrogen active radicals. For instance, it was
shown that NaPF/EC–DEC-based electrolytes containing wt%
ethoxy(pentafluoro)cyclotriphosphazene (EFPN) can eectively
reduce the self-extinguishing time from  to s, demon-
strating the successful synthesis of nonflammable electrolytes
(Figures e,f).[] However, it should be noted that introducing
EFPN with a low dielectric constant reduces the ionic conductiv-
ity of the resulting electrolyte. Moreover, other flame-retardant
additives, such as perfluoro--methyl--pentanone, trimethyl
phosphate (TMP), triethyl phosphate, triphenyl phosphate, di-
ethyl ethylphosphonate, and cresyl diphenyl phosphate, have
been developed, but research on flame-retardant additives in
SIBs systems is still limited.[, ] Recently, high-concentration
electrolytes have attracted attention in both LIB and SIB
systems owing to their unique electrochemical properties,
Adv. Energy Mater. 2023, 2301975 © 2023 Wiley-VCH GmbH
2301975 (32 of 41)
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202301975 by Hanyang University Library, Wiley Online Library on [18/08/2023]. 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
www.advancedsciencenews.com www.advenergymat.de
including wide voltage windows, enhanced thermal stability, and
inhibition of dendrite growth.[ ] However, the high flame re-
tardancy of high-concentration electrolytes is related to the low
content of flammable organic solvents, so high-concentration
electrolytes without flame-retardant additives may still ignite.
In addition, the intrinsic drawbacks of high-concentration elec-
trolytes, such as poor wettability toward the electrode/separator,
high viscosity, and high cost, cannot be ignored.[] To balance
the ionic conductivity, viscosity, and electrochemical/thermal
stability, researchers introduced inert and poorly solvating sol-
vents (fluorinated ether (BTFE)) into the high-concentration elec-
trolyte to obtain a localized high-concentration electrolyte (.
NaTFSI in TMP/BTFE/VC).[] The potential of this non-
flammable electrolyte was proven by the electrochemical and
safety performance of practical pouch cells. In fact, high pro-
duction costs are the main bottleneck for the practical appli-
cation of high-concentration/localized high-concentration elec-
trolytes. Ionic liquids with high electrochemical/thermal stabil-
ity and non-flammability are expected to enable the development
of highly safe SIBs, but they need to overcome the current lim-
itations of high cost, high viscosity, and low conductivity.[ ] To
date, the optimization strategies for ionic liquids have focused
on the development of hybrid electrolytes, such as those con-
taining gel and organic electrolytes. As the “Holy Grail” of high-
energy-density battery systems, the development of all-solid-state
SIBs has also received significant attention.[ ] Because solid-
state electrolytes used in SIBs are still in the research phase, some
key challenges need a major breakthrough, such as the low ionic
conductivity, high solid–solid interfacial resistance, and diculty
in manufacturing solid-state batteries. However, these issues are
beyond the scope of this review and will not be discussed further.
The conventional electrolytes for SIBs that can match with
high-performance cathode materials are mainly binary (PC +
FEC) or ternary (EC +DEC +FEC, EC +PC +FEC, EC +
DMC +FEC) solvents with мNaPFor NaClO. Among them,
the PB cathode shows similar electrochemical performance in
NaPF-andNaClO
-based electrolytes, which may be related to
the low water sensitivity of NaClO.[] On the contrary, the CEI
layer formed by the decomposition of NaPFexhibits higher Na-
conductivity than NaClO, resulting in lower interfacial resis-
tance and higher cyclic stability.[] In addition, the dissolved
SEI species produced by linear carbonates (DMC, EMC, DEC)
reduction shuttle from the anode to the cathode side for oxida-
tion, resulting in Na loss and capacity decay.[  ] Therefore, the
electrolyte composed of NaPFand single cyclic carbonates (PC
and EC) is more popular in layered oxides and polyanionic com-
pounds. As the most eective electrolyte additive, introducing
FEC into the electrolyte can induce the formation of a robust F-
rich CEI layer on the surface of cathode material, thereby improv-
ing cycle stability. However, it should be noted that FEC is not a
perfect electrolyte additive as it does not match with some an-
odes and is prone to decomposition at elevated temperatures.[ ]
On the other hand, although fluorinated electrolyte components
exhibit high electrochemical stability, the production of toxic
byproducts (HF, NaF) can pose safety and environmental haz-
ards. Therefore, it is suggested to develop novel functional elec-
trolytes that are inexpensive, non-toxic, and can form thin and
uniform interface layers on the cathode and anode, so as to fur-
ther promote the development of SIBs. In summary, optimiz-
ing cathode materials and matching appropriate anodes and elec-
trolytes are the key to realize commercial application of SIBs.
5. Conclusions and Perspectives
SIBs with cost and safety advantages have shown great potential
for replacing LIBs in next-generation large-scale energy-storage
systems and moderate-range electric vehicles. In full-cell appli-
cations, O-type layered oxides with sucient Na are considered
the most promising cathode candidates owing to their high oper-
ating voltage, high reversible capacity, long service life, low pro-
duction cost, and practicality. However, their sluggish Na dynam-
ics, complex phase-evolution processes, tendency to form micro-
cracks and undergo surface reconstruction during cycling, and
air sensitivity are the main factors limiting their large-scale ap-
plication. Recent literature shows that customizing layered ox-
ides by rational compositional/structural design can eectively
overcome some of the inherent shortcomings, which is of great
significance for promoting the commercialization of SIBs. There-
fore, this review summarizes the current development status of
O-type cathodes in SIBs and provides constructive prospects for
their large-scale applications.
) Appropriate compositional design not only improves the
structural stability/reversibility of layered oxides through the
synergistic eect of multiple TM ions, but also fundamentally
overcomes the capacity loss caused by structural degradation
by smoothing the phase evolution process or extending the
solid-solution reaction zone. In addition, coupling metal ions
with high redox reaction pairs (such as Cu+/+) or with dif-
ferent cationic potential (such as Ti and Mg) in the TMO
slab can increase the energy density by increasing the aver-
age working potential, or improve the air stability by adjust-
ing the interlayer distance. In recent years, high-entropy ox-
ides and multiphase cathodes synthesized based on compo-
sitional design have also received research attention owing to
their unique electrochemical performance, but their practical
application potential needs further verification.
) Surface-passivation layers provide a physical barrier to pre-
vent direct contact between the active material and corrosive
electrolytes and/or humid air, thereby suppressing parasitic
reactions and improving interfacial and air stability. These
surface-modified layers act as buer matrices to alleviate the
volume expansion and dissolution of active materials dur-
ing the electrochemical process. In addition, the combina-
tion of surface coating and bulk doping can enhance the sur-
face/structural stability and reduce ionic-diusion barriers,
thus further improving the cycling stability and rate proper-
ties. Moreover, the consumption of surface residual Na and
the in-situ formation of Na-containing layers (for example,
phosphate and P phase layers) can further improve the Na
storage performance and air/thermal stability of layered ox-
ides.
) Optimizing the synthesis methods to design cathode parti-
cles with unique morphology, including microspheres, fibers,
flakes, and single-crystal structures, is beneficial for shorten-
ing the Na-ion diusion paths, exposing active planes, and ac-
commodating internal stress. Among them, spherical materi-
als have high tap density and low specific surface area to meet
Adv. Energy Mater. 2023, 2301975 © 2023 Wiley-VCH GmbH
2301975 (33 of 41)
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202301975 by Hanyang University Library, Wiley Online Library on [18/08/2023]. 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
www.advancedsciencenews.com www.advenergymat.de
the requirements of high volume energy density and minimal
surface side reactions, providing insights to support the large-
scale production of SIBs.
Although the Na-ion storage performance of O-type cathode
materials has been greatly improved after years of eort, the fol-
lowing topics require further attention.
) Balancing production costs and electrochemical performance.For
various layered oxide candidates, it is recommended to fur-
ther develop binary or ternary cathode materials contain-
ing Ni (low/medium content), Mn (medium content), Fe
(low/medium content), and Cu (low content). In addition, it
has been proven that metal-ion substitution can further im-
prove the electrochemical performance of oxide materials, but
the doping content should not be high enough to risk sacrific-
ing the reversible capacity (especially for high-entropy oxides).
Furthermore, it is necessary to avoid introducing toxic and/or
expensive metals (such as Li, Sn, and Sb) that will weaken
the competitiveness of SIBs. Surface modification can protect
air-sensitive layered oxides by inhibiting structural degrada-
tion, but it also increases the number of production steps and
costs. Therefore, a one-step integrated doping–coating syn-
thesis method for cathode materials could provide a promis-
ing modification strategy.
) Expand production capacity and batch stability.Themain
method used for the large-scale production of cathode ma-
terials is the fabrication of uniform spherical secondary par-
ticles by a coprecipitation method, which has been demon-
strated for layered LIBs. Therefore, subsequent studies are
required to optimize the synthesis process to develop parti-
cles with unique structures, such as core–shell and radiation
structures, and exposed active planes.
) Evaluate practicality using pouch-type Na-ion full cells. A prac-
tical Na full-cell system requires combining a cathode mate-
rial (with excellent electrochemical performance) with a suit-
able HC anode material (to provide a high ICE) and a non-
flammable electrolyte (to enable a wide voltage window). In
addition, it should be noted that the unit energy cost of SIBs
is closely related to the production cost of each component
and the electrolyte usage. More importantly, the practicality
of SIBs needs to be evaluated using pouch-type cells to clar-
ify the potential operating temperature range, gas production
issues, and overall safety.
) Investigate the mechanism of capacity fading. It is necessary
to further investigate the capacity fading caused by struc-
tural degradation and side reactions to guide the develop-
ment of SIBs. Therefore, it is necessary to combine ad-
vanced characterization techniques such as neutron dirac-
tion, synchrotron X-ray diraction, spherical aberration-
corrected scanning transmission electron microscopy, nu-
clear magnetic resonance, X-ray absorption spectroscopy,
and theoretical calculations to provide a deep understand-
ing of the relationships between the crystal structure, chem-
istry/electrochemistry properties, and Na-ion storage perfor-
mance of O-type cathode materials.
) Develop next-generation advanced cathode materials. SIBs that
can be used in electric vehicles require fast charging, so it is
necessary to further optimize the cathode materials to achieve
high power density and alleviate the long charging times of
current electric vehicles. In addition, the reversible capacity
of cathode materials can be further improved by inducing
anionic redox reactions through doping and defect forma-
tion. However, because anionic redox reactions usually oc-
cur under high voltage (. V), which may lead to O-loss and
structural degradation, it is necessary to conduct in-depth re-
search on the reaction mechanism and investigate possible
solutions.
By combining multiple optimization strategies,[] O-type
cathodes have achieved high capacity (> mA h g), high op-
erating voltage (>. V), and high thermal stability (exothermal
peak > °C) in SIB systems, which are now considered a vi-
able competitor to commercial LIBs.[ ] In the coming years, we
hope that the electrochemical performance of SIBs will be further
improved, creating opportunities in both academia and industry.
Acknowledgements
X.L. and J.-Y.H. contributed equally to this work. This work was supported
by the Human Resources Development Program (No. 20214000000320)
of the Korea Institute of Energy Technology Evaluation and Planning
(KETEP), funded by the Ministry of Trade, Industry, and Energy of the Ko-
rean government. The authors acknowledge Hoon-Hee Ryu and Gwan-
geon Oh for useful discussions.
Conflict of Interest
The authors declare no conflict of interest.
Keywords
cathode materials, commercialization, O3-type structures, sodium ion
batteries, strategies
Received: June 23, 2023
Revised: August 1, 2023
Published online:
[1] N. Voronina, Y.-K. Sun, S.-T. Myung, ACS Energy Lett. 2020,5, 1814.
[2] J.-Y. Hwang, S.-T. Myung, Y.-K. Sun, Chem.Soc.Rev.2017,46, 3529.
[3] A. Rudola, R. Sayers, C. J. Wright, J. Barker, Nat. Energy 2023,8, 215.
[4] Y. Kim, K.-H. Ha, S. M. Oh, K. T. Lee, Chemistry 2014,20, 11980.
[5] J. W. Choi, D. Aurbach, Nat. Rev. Mater. 2016,1, 16013.
[6] E. Goikolea, V. Palomares, S. Wang, I. R. Larramendi, X. Guo, G.
Wang, T. Rojo, Adv. Energy Mater. 2020,10, 2002055.
[7] R. Usiskin, Y. Lu, J. Popovic, M. Law, P. Balaya, Y.-S. Hu, J. Maier,
Nat. Rev. Mater. 2021,6, 1020.
[8] A.Rudola,A.J.R.Rennie,R.Heap,S.S.Meysami,A.Lowbridge,F.
Mazzali, R. Sayers, C. J. Wright, J. Barker, J. Mater. Chem. A 2021,9,
8279.
[9] K.Chayambuka,G.Mulder,D.L.Danilov,P.H.L.Notten,Adv. En-
ergy Mater. 2020,10, 2001310.
[10] K. Wu, X. Dou, X. Zhang, C. Ouyang, Engineering 2023,21, 36.
[11] A. Bauer, J. Song, S. Vail, W. Pan, J. Barker, Y. Lu, Adv. Energy Mater.
2018,8, 1702869.
[12] A. Fukunaga, T. Nohira, R. Hagiwara, K. Numata, E. Itani, S. Sakai,
K. Nitta, J. Appl. Electrochem. 2016,46, 487.
Adv. Energy Mater. 2023, 2301975 © 2023 Wiley-VCH GmbH
2301975 (34 of 41)
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202301975 by Hanyang University Library, Wiley Online Library on [18/08/2023]. 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
www.advancedsciencenews.com www.advenergymat.de
[13] A. Tripathi, A. Rudola, S. R. Gajjela, S. Xi, P. Balaya, J. Mater. Chem.
A2019,7, 25944.
[14] S. Mariyappan, T. Marchandier, F. Rabuel, A. Iadecola, G. Rousse,
A. V. Morozov, A. M. Abakumov, J.-M. Tarascon, Chem. Mater. 2020,
32, 1657.
[15] Y. Li, Q. Zhou, S. Weng, F. Ding, X. Qi, J. Lu, Y. Li, X. Zhang, X. Rong,
Y. Lu, X. Wang, R. Xiao, H. Li, X. Huang, L. Chen, Y.-S. Hu, Nat.
Energy 2022,7, 511.
[16] Q. Liu, Z. Hu, M. Chen, C. Zou, H. Jin, S. Wang, S. L. Chou, Y. Liu,
S. X. Dou, Adv. Funct. Mater. 2020,30, 1909530.
[17] X. Xiang, K. Zhang, J. Chen, Adv. Mater. 2015,27, 5343.
[18] H. Li, M. Xu, Z. Zhang, Y. Lai, J. Ma, Adv. Funct. Mater. 2020,30,
2000473.
[19] T. Jin, H. Li, K. Zhu, P.-F. Wang, P. Liu, L. Jiao, Chem.Soc.Rev.2020,
49, 2342.
[20] L. Liang, X. Li, F. Zhao, J. Zhang, Y. Liu, L. Hou, C. Yuan, Adv. Energy
Mater. 2021,11, 2100287.
[21] X. Liang, X. Ou, F. Zheng, Q. Pan, X. Xiong, R. Hu, C. Yang, M. Liu,
ACS Appl. Mater. Interfaces 2017,9, 13151.
[22] Q. Wei, X. Chang, J. Wang, T. Huang, X. Huang, J. Yu, H. Zheng, J.
H. Chen, D. L. Peng, Adv. Mater. 2022,34, 2108304.
[23] L. Shen, Y. Li, S. Roy, X. Yin, W. Liu, S. Shi, X. Wang, X. Yin, J. Zhang,
Y. Z h a o , Chin. Chem. Lett. 2021,32, 3570.
[24] X. Cao, A. Pan, B. Yin, G. Fang, Y. Wang, X. Kong, T. Zhu, J. Zhou, G.
Cao, S. Liang, Nano Energy 2019,60, 312.
[25] T. Tong, Z. Tian, W. Chen, Y. Linghu, D. Li, Z. Tian, Y. Wang, Y. Chen,
L. Guo, Electrochim. Acta 2022,411, 140073.
[26] J. Barker, M. Y. Saidi, J. L. Swoyer, Electrochem. Solid-State Lett. 2003,
6,A1.
[27] X.Ge,X.Li,Z.Wang,H.Guo,G.Yan,X.Wu,J.Wang,Chem. Eng. J.
2019,357, 458.
[28] M. Ling, Q. Jiang, T. Li, C. Wang, Z. Lv, H. Zhang, Q. Zheng, X. Li,
Adv. Energy Mater. 2021,11, 2100627.
[29] L. Li, Y. Xu, X. Sun, R. Chang, Y. Zhang, X. Zhang, J. Li, Adv. Energy
Mater. 2018,8, 1801064.
[30] M. Bianchini, N. Brisset, F. Fauth, F. Weill, E. Elkaim, E. Suard, C.
Masquelier, L. Croguennec, Chem. Mater. 2014,26, 4238.
[31] K. Liang, H. Zhao, J. Li, X. Huang, S. Jia, W. Chen, Y. Ren, Small 2023,
19, 2207562.
[32] Y. Li, X. Liang, G. Chen, W. Zhong, Q. Deng, F. Zheng, C. Yang, M.
Liu, J. Hu, Chem. Eng. J. 2020,387, 123952.
[33] Z. Y. Gu, J. Z. Guo, J. M. Cao, X. T. Wang, X. X. Zhao, X. Y. Zheng, W.
H.Li,Z.H.Sun,H.J.Liang,X.L.Wu,Adv. Mater. 2022,34, 2110108.
[34] W. Zhou, L. Xue, X. Lü, H. Gao, Y. Li, S. Xin, G. Fu, Z. Cui, Y. Zhu, J.
B. Goodenough, Nano Lett. 2016,16, 7836.
[35] H. Li, T. Jin, X. Chen, Y. Lai, Z. Zhang, W. Bao, L. Jiao, Adv. Energy
Mater. 2018,8, 1801418.
[36] L. Zhu, M. Zhang, L. Yang, K. Zhou, Y. Wang, D. Sun, Y. Tang, H.
Wan g, Nano Energy 2022,99, 107396.
[37] Y. Wu, X. Meng, L. Yan, Q. Kang, H. Du, C. Wan, M. Fan, T. Ma, J.
Mater. Chem. A 2022,10, 21816.
[38] H. Gao, Y. Li, K. Park, J. B. Goodenough, Chem. Mater. 2016,28,
6553.
[39] T. Zhu, P. Hu, C. Cai, Z. Liu, G. Hu, Q. Kuang, L. Mai, L. Zhou, Nano
Energy 2020,70, 104548.
[40] H. Li, M. Xu, C. Gao, W. Zhang, Z. Zhang, Y. Lai, L. Jiao, Energy Stor-
age Mater. 2020,26, 325.
[41] P. Barpanda, T. Ye, S.-I. Nishimura, S.-C. Chung, Y. Yamada, M.
Okubo, H. Zhou, A. Yamada, Electrochem. Commun. 2012,24,
116.
[42] P. Barpanda, G. Liu, C. D. Ling, M. Tamaru, M. Avdeev, S.-C. Chung,
Y. Yamada, A. Yamada, Chem. Mater. 2013,25, 3480.
[43] C.S.Park,H.Kim,R.A.Shakoor,E.Yang,S.Y.Lim,R.Kahraman,
Y. Jung, J. W. Choi, J. Am. Chem. Soc. 2013,135, 2787.
[44] H. Kim, G. Yoon, I. Park, K.-Y. Park, B. Lee, J. Kim, Y.-U. Park, S.-K.
Jung, H.-D. Lim, D. Ahn, S. Lee, K. Kang, Energy Environ. Sci. 2015,
8, 3325.
[45] H. Kim, I. Park, D.-H. Seo, S. Lee, S.-W. Kim, W. J. Kwon, Y.-U. Park,
C. S. Kim, S. Jeon, K. Kang, J. Am. Chem. Soc. 2012,134, 10369.
[46] A. Paolella, C. Faure, V. Timoshevskii, S. Marras, G. Bertoni, A.
Guerfi, A. Vijh, M. Armand, K. Zaghib, J. Mater. Chem. A 2017,5,
18919.
[47] W. Ren, Z. Zhu, M. Qin, S. Chen, X. Yao, Q. Li, X. Xu, Q. Wei, L. Mai,
C. Zhao, Adv. Funct. Mater. 2019,29, 1806405.
[48] J. Qian, C. Wu, Y. Cao, Z. Ma, Y. Huang, X. Ai, H. Yang, Adv. Energy
Mater. 2018,8, 1702619.
[49] J. Peng, M. Ou, H. Yi, X. Sun, Y. Zhang, B. Zhang, Y. Ding, F. Wang,
S. Gu, C. A. López, W. Zhang, Y. Liu, J. Fang, P. Wei, Y. Li, L. Miao, J.
Jiang, C. Fang, Q. Li, M. T. Fernández-Díaz, J. A. Alonso, S. Chou, J.
Han, Energy Environ. Sci. 2021,14, 3130.
[50] L. Wang, J. Song, R. Qiao, L. A. Wray, M. A. Hossain, Y.-D. Chuang,
W. Yang, Y. Lu, D. Evans, J.-J. Lee, S. Vail, X. Zhao, M. Nishijima, S.
Kakimoto, J. B. Goodenough, J. Am. Chem. Soc. 2015,137, 2548.
[51] Y. Tang, W. Li, P. Feng, M. Zhou, K. Wang, Y. Wang, K. Zaghib, K.
Jiang, Adv. Funct. Mater. 2020,30, 1908754.
[52] B. Xie, B. Sun, T. Gao, Y. Ma, G. Yin, P. Zuo, Coord. Chem. Rev. 2022,
460, 214478.
[53] H. Zhang, Y. Gao, X. Liu, L. Zhou, J. Li, Y. Xiao, J. Peng, J. Wang, S.
L. Chou, Adv. Energy Mater. 2023,13, 2300149.
[54] H. Zhang, J. Peng, L. Li, Y. Zhao, Y. Gao, J. Wang, Y. Cao, S. Dou, S.
Chou, Adv. Funct. Mater. 2023,33, 2210725.
[55] L.-L.Zhang,Z.-Y.Chen,X.-Y.Fu,B.Yan,H.-C.Tao,X.-L.Yang,Chem.
Eng. J. 2022,433, 133739.
[56] Z. Xu, Y. Sun, J. Xie, Y. Nie, X. Xu, J. Tu, C. Shen, Y. Jin, Y. Li, Y. Lu,
A. Zhou, F. Chen, T. Zhu, X. Zhao, Mater. Today: Sustain. 2022,18,
100113.
[57] Y. Ma, Y. Ma, S. L. Dreyer, Q. Wang, K. Wang, D. Goonetilleke, A.
Omar, D. Mikhailova, H. Hahn, B. Breitung, T. Brezesinski, Adv.
Mater. 2021,33, 2101342.
[58] J. Peng, Y. Gao, H. Zhang, Z. Liu, W. Zhang, L. Li, Y. Qiao, W. Yang, J.
Wang,S.Dou,S.Chou,Angew. Chem., Int. Ed. 2022,61, e202205867.
[59] W.Wang,Y.Gang,J.Peng,Z.Hu,Z.Yan,W.Lai,Y.Zhu,D.Appadoo,
M. Ye, Y. Cao, Q. F. Gu, H. K. Liu, S. X. Dou, S. L. Chou, Adv. Funct.
Mater. 2022,32, 2111727.
[60] J. Hu, H. Tao, M. Chen, Z. Zhang, S. Cao, Y. Shen, K. Jiang, M. Zhou,
ACS Appl. Mater. Interfaces 2022,14, 12234.
[61] J. Song, L. Wang, Y. Lu, J. Liu, B. Guo, P. Xiao, J. J. Lee, X. Q. Yang,
G. Henkelman, J. B. Goodenough, J. Am. Chem. Soc. 2015,137,
2658.
[62] W.-J. Li, S.-L. Chou, J.-Z. Wang, Y.-M. Kang, J.-L. Wang, Y. Liu, Q.-F.
Gu, H.-K. Liu, S.-X. Dou, Chem. Mater. 2015,27, 1997.
[63] M. Qin, W. Ren, R. Jiang, Q. Li, X. Yao, S. Wang, Y. You, L. Mai, ACS
Appl. Mater. Interfaces 2021,13, 3999.
[64] Z.-E. Yu, H. Cheng, Y. Lyu, Y. Liu, J. Zhou, R. Chen, B. Guo, J. Alloys
Compd. 2021,887, 161388.
[65] J. Peng, W. Zhang, Z. Hu, L. Zhao, C. Wu, G. Peleckis, Q. Gu, J. Z.
Wang,H.K.Liu,S.X.Dou,S.Chou,Nano Lett. 2022,22, 1302.
[66] M. Jiang, Z. Hou, J. Wang, L. Ren, Y. Zhang, J.-G. Wang, Nano Energy
2022,102, 107708.
[67] S.Wang,M.Qin,M.Huang,X.Huang,Q.Li,Y.You,ACS Appl. En-
ergy Mater. 2022,5, 6927.
[68] C. Duan, Y. Meng, Y. Wang, Z. Zhang, Y. Ge, X. Li, Y. Guo, D. Xiao,
Inorg. Chem. Front. 2021,8, 2008.
[69] G. He, L. F. Nazar, ACS Energy Lett. 2017,2, 1122.
[70] Y. Huang, X. Zhang, L. Ji, L. Wang, B. B. Xu, M. W. Shahzad, Y. Tang,
Y. Zhu, M. Yan, G. Sun, Y. Jiang, Energy Storage Mater. 2023,58,1.
[71] D. Yang, J. Xu, X. Z. Liao, H. Wang, Y. S. He, Z. F. Ma, Chem. Com-
mun. 2022,59, 211.
Adv. Energy Mater. 2023, 2301975 © 2023 Wiley-VCH GmbH
2301975 (35 of 41)
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202301975 by Hanyang University Library, Wiley Online Library on [18/08/2023]. 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
www.advancedsciencenews.com www.advenergymat.de
[72] Y.Xiao,N.M.Abbasi,Y.F.Zhu,S.Li,S.J.Tan,W.Ling,L.Peng,T.
Yang, L. Wang, X. D. Guo, Y. X. Yin, H. Zhang, Y. G. Guo, Adv. Funct.
Mater. 2020,30, 2001334.
[73] Y.Liu,D.Wang,H.Li,P.Li,Y.Sun,Y.Liu,Y.Liu,B.Zhong,Z.Wu,X.
Guo, J. Mater. Chem. A 2022,10, 3869.
[74] Q. Liu, Z. Hu, M. Chen, C. Zou, H. Jin, S. Wang, Q. Gu, S. Chou, J.
Mater. Chem. A 2019,7, 9215.
[75] Y.-K. Sun, ACS Energy Lett. 2020,5, 1278.
[76] E. Talaie, V. Duffort, H. L. Smith, B. Fultz, L. F. Nazar, Energy Environ.
Sci. 2015,8, 2512.
[77] S. Mariyappan, Q. Wang, J. M. Tarascon, J. Electrochem. Soc. 2018,
165, A3714.
[78] P. Kulkarni, H. Jung, D. Ghosh, M. Jalalah, M. Alsaiari, F. A. Harraz,
R. G. Balakrishna, J Energy Chem 2023,76, 479.
[79] H. S. Hirsh, Y. Li, D. H. S. Tan, M. Zhang, E. Zhao, Y. S. Meng, Adv.
Energy Mater. 2020,10, 2001274.
[80] W. J. Li, C. Han, W. Wang, F. Gebert, S. L. Chou, H. K. Liu, X. Zhang,
S. X. Dou, Adv. Energy Mater. 2017,7, 1700274.
[81] W. Zuo, A. Innocenti, M. Zarrabeitia, D. Bresser, Y. Yang, S.
Passerini, Acc. Chem. Res. 2023,56, 284.
[82] N. Tapia-Ruiz, A. R. Armstrong, H. Alptekin, M. A. Amores, H. Au, J.
Barker, R. Boston, W. R. Brant, J. M. Brittain, Y. Chen, M. Chhowalla,
Y.-S. Choi, S. I. R. Costa, M. Crespo Ribadeneyra, S. A. Cussen, E. J.
Cussen, W. I. F. David, A. V. Desai, S. A. M. Dickson, E. I. Eweka, J.
D. Forero-Saboya, C. P. Grey, J. M. Griffin, P. Gross, X. Hua, J. T. S.
Irvine, P. Johansson, M. O. Jones, M. Karlsmo, E. Kendrick, et al.,
JPhys Energy 2021,3, 031503.
[83] L.Li,G.Tan,J.Tao,Z.Lun,C.Xu,ACS Appl. Energy Mater. 2023,6,
6883.
[84] X. Liang, Y.-K. Sun, Adv. Funct. Mater. 2022,32, 2206154.
[85] Z. Li, M. Dadsetan, J. Gao, S. Zhang, L. Cai,A. Naseri, M. E. Jimenez-
Castaneda, T. Filley, J. T. Miller, M. J. Thomson, V. G. Pol, Adv. Energy
Mater. 2021,11, 2101764.
[86] A. J. Toumar, S. P. Ong, W. D. Richards, S. Dacek, G. Ceder, Phys.
Rev. Appl. 2015,4, 064002.
[87] X. Li, X. Ma, D. Su, L. Liu, R. Chisnell, S. P. Ong, H. Chen, A. Toumar,
J. C. Idrobo, Y. Lei, J. Bai, F. Wang, J. W. Lynn, Y. S. Lee, G. Ceder, Nat.
Mater. 2014,13, 586.
[88] N. A. Katcho, J. Carrasco, D. Saurel, E. Gonzalo, M. Han, F. Aguesse,
T. Rojo, Adv. Energy Mater. 2017,7, 1601477.
[89] M. D. Radin, J. Alvarado, Y. S. Meng, A. Van Der Ven, Nano Lett.
2017,17, 7789.
[90] K. Kubota, T. Asari, H. Yoshida, N. Yaabuuchi, H. Shiiba, M.
Nakayama, S. Komaba, Adv. Funct. Mater. 2016,26, 6047.
[91] Q. Huang, Y. Feng, L. Wang, S. Qi, P. He, X. Ji, C. Liang, S. Chen, L.
Zhou, W. Wei, Chem. Eng. J. 2022,431, 133454.
[92] F. Ding, C. Zhao, D. Zhou, Q. Meng, D. Xiao, Q. Zhang, Y. Niu, Y.
Li, X. Rong, Y. Lu, L. Chen, Y.-S. Hu, Energy Storage Mater. 2020,30,
420.
[93] J. Su, Y. Pei, Z. Yang, X. Wang, RSC Adv. 2015,5, 27229.
[94] S. Kim, X. Ma, S. P. Ong, G. Ceder, Phys.Chem.Chem.Phys.2012,
14, 15571.
[95] M. Tang, J. Yang, H. Liu, X. Chen, L. Kong, Z. Xu, J. Huang, Y. Xia,
ACS Appl. Mater. Interfaces 2020,12, 45997.
[96] S. P. Ong, V. L. Chevrier, G. Hautier, A. Jain, C. Moore, S. Kim, X.
Ma, G. Ceder, Energy Environ. Sci. 2011,4, 3680.
[97] N. Yabuuchi, S. Komaba, Sci. Technol. Adv. Mater. 2014,15, 043501.
[98] S. Liu, B. Wang, X. Zhang, S. Zhao, Z. Zhang, H. Yu, Matter 2021,4,
1511.
[99] P.-F. Wang, T. Jin, J. Zhang, Q.-C. Wang, X. Ji, C. Cui, N. Piao, S. Liu,
J. Xu, X.-Q. Yang, C. Wang, Nano Energy 2020,77, 105167.
[100] H. Jung, J. Kim, S. Kim, J. Appl. Phys. 2022,132, 055101.
[101] P.-F. Wang, Y. You, Y.-X. Yin, Y.-G. Guo, J. Mater. Chem. A 2016,4,
17660.
[102] Y. Xie, H. Wang, G. Xu, J. Wang, H. Sheng, Z. Chen, Y. Ren, C.-J. Sun,
J. Wen, J. Wang, D. J. Miller, J. Lu, K. Amine, Z.-F. Ma, Adv. Energy
Mater. 2016,6, 1601306.
[103] S. Komaba, N. Yabuuchi, T. Nakayama, A. Ogata, T. Ishikawa, I.
Nakai, Inorg. Chem. 2012,51, 6211.
[104] M. H. Han, E. Gonzalo, M. Casas-Cabanas, T. Rojo, J. Power Sources
2014,258, 266.
[105] X. Li, D. Wu, Y.-N. Zhou, L. Liu, X.-Q. Yang, G. Ceder, Electrochem.
Commun. 2014,49, 51.
[106] B. Xiao, Y. Wang, S. Tan, M. Song, X. Li, Y. Zhang, F. Lin, K. S. Han,
F. Omenya, K. Amine, X. Q. Yang, D. Reed, Y. Hu, G. L. Xu, E. Hu, X.
Li, X. Li, Angew. Chem., Int. Ed. 2021,60, 8258.
[107] T.-Y. Yu, H.-H. Ryu, G. Han, Y.-K. Sun, Adv. Energy Mater. 2020,10,
2001609.
[108] S. Lee, S. W. Doo, M. S. Jung, S. G. Lim, K. Kim, K. T. Lee, J. Mater.
Chem. A 2021,9, 14074.
[109] L. Mu, X. Feng, R. Kou, Y. Zhang, H. Guo, C. Tian, C. J. Sun, X. W. Du,
D. Nordlund, H. L. Xin, F. Lin, Adv. Energy Mater. 2018,8, 1801975.
[110] H.-H. Ryu, K.-J. Park, C. S. Yoon, Y.-K. Sun, Chem. Mater. 2018,30,
1155.
[111] Y. Li, X. Li, C. Du, H. Sun, Y. Zhang, Q. Liu, T. Yang, J. Zhao, C.
Delmas, S. J. Harris, H. Chen, Q. Huang, Y. Tang, L. Zhang, T. Zhu,
J. Huang, ACS Energy Lett. 2021,6, 3960.
[112] J. Lamb, L. Stokes, A. Manthiram, Chem. Mater. 2020,32, 7389.
[113] L. Gan, X. G. Yuan, J. J. Han, X. Yang, L. Zheng, Z. Huang, H. R. Yao,
Adv. Funct. Mater. 2023,33, 2209026.
[114] T.-Y. Yu, J. Kim, J.-Y. Hwang, H. Kim, G. Han, H.-G. Jung, Y.-K. Sun,
J. Mater. Chem. A 2020,8, 13776.
[115] W. Zuo, J. Qiu, X. Liu, F. Ren, H. Liu, H. He, C. Luo, J. Li, G. F. Ortiz,
H. Duan, J. Liu, M.-S. Wang, Y. Li, R. Fu, Y. Yang, Nat. Commun. 2020,
11, 3544.
[116] S. Roberts, L. Chen, B. Kishore, C. E. J. Dancer, M. Simmons, E.
Kendrick, J. Colloid Interface Sci. 2022,627, 427.
[117] L. Zheng, L. Li, R. Shunmugasundaram, M. N. Obrovac, ACS Appl.
Mater. Interfaces 2018,10, 38246.
[118] J. Lamb, A. Manthiram, ACS Appl. Energy Mater. 2021,4, 11735.
[119] X. G. Yuan, Y. J. Guo, L. Gan, X. A. Yang, W. H. He, X. S. Zhang, Y. X.
Yin, S. Xin, H. R. Yao, Z. Huang, Y. G. Guo, Adv. Funct. Mater. 2022,
32, 2111466.
[120] T. Song, E. Kendrick, JPhys Mater. 2021,4, 032004.
[121] S.-H. Bo, X. Li, A. J. Toumar, G. Ceder, Chem. Mater. 2016,28,
1419.
[122] C. Chen, Z. Ding, Z. Han, C. Liang, X. Lan, P. Wang, P. Gao, W. Wei,
J. Phys. Chem. Lett. 2020,11, 5464.
[123] D. Susanto, M. K. Cho, G. Ali, J.-Y. Kim, H. J. Chang, H.-S. Kim, K.-W.
Nam, K. Y. Chung, Chem. Mater. 2019,31, 3644.
[124] B. Silván, E. Gonzalo, L. Djuandhi, N. Sharma, F. Fauth, D. Saurel,
J. Mater. Chem. A 2018,6, 15132.
[125] X. Li, Y. Wang, D. Wu, L. Liu, S.-H. Bo, G. Ceder, Chem. Mater. 2016,
28, 6575.
[126] X. Gao, L. Fang, H. Wang, S. Lee, H. Liu, S. Zhang, J. Gao, Y. Mei,
M. Park, J. Zhang, M. Chen, L. Zhou, W. Deng, G. Zou, H. Hou, Y.
M. Kang, X. Ji, Adv. Funct. Mater. 2023,33, 2212685.
[127] S. Chu, D. Kim, G. Choi, C. Zhang, H. Li, W. K. Pang, Y. Fan, A.
M. D’angelo, S. Guo, H. Zhou, Angew. Chem., Int. Ed. 2023,135,
e202216174.
[128] X. Hou, X. Liu, H. Wang, X. Zhang, J. Zhou, M. Wang, Energy Storage
Mater. 2023,57, 577.
[129] H. Xu, Z. Li, T. Liu, C. Han, C. Guo, H. Zhao, Q. Li, J. Lu, K. Amine,
X. Qiu, Angew. Chem., Int. Ed. 2022,61, e202202894.
[130] E. Lee, D. E. Brown, E. E. Alp, Y. Ren, J. Lu, J.-J. Woo, C. S. Johnson,
Chem. Mater. 2015,27, 6755.
[131] S. Komaba, C. Takei, T. Nakayama, A. Ogata, N. Yabuuchi, Elec-
trochem. Commun. 2010,12, 355.
Adv. Energy Mater. 2023, 2301975 © 2023 Wiley-VCH GmbH
2301975 (36 of 41)
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202301975 by Hanyang University Library, Wiley Online Library on [18/08/2023]. 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
www.advancedsciencenews.com www.advenergymat.de
[132] L. Mu, S. Xu, Y. Li, Y. S. Hu, H. Li, L. Chen, X. Huang, Adv. Mater.
2015,27, 6928.
[133] K. Wang, H. Zhuo, J. Wang, F. Poon, X. Sun, B. Xiao, Adv. Funct.
Mater. 2023,33, 2212607.
[134] S. Chu, S. Guo, H. Zhou, Chem.Soc.Rev.2021,50, 13189.
[135] I. Lee, G. Oh, S. Lee, T.-Y. Yu, M. H. Alfaruqi, V. Mathew, B.
Sambandam, Y.-K. Sun, J.-Y. Hwang, J. Kim, Energy Storage Mater.
2021,41, 183.
[136] C.-Y. Yu, J.-S. Park, H.-G. Jung, K.-Y. Chung, D. Aurbach, Y.-K. Sun,
S.-T. Myung, Energy Environ. Sci. 2015,8, 2019.
[137] Y. Wang, W. Li, G. Hu, Z. Peng, Y. Cao, H. Gao, K. Du, J. B.
Goodenough, Chem. Mater. 2019,31, 5214.
[138] J.-J. Ding, Y.-N. Zhou, Q. Sun, Z.-W. Fu, Electrochem. Commun. 2012,
22, 85.
[139] L. Liang, X. Sun, D. K. Denis, J. Zhang, L. Hou, Y. Liu, C. Yuan, ACS
Appl. Mater. Interfaces 2019,11, 4037.
[140] W. Ko, M.-K. Cho, J. Kang, H. Park, J. Ahn, Y. Lee, S. Lee, S. Lee, K.
Heo, J. Hong, J.-K. Yoo, J. Kim, Energy Storage Mater. 2022,46, 289.
[141] P. F. Wang, Y. You, Y. X. Yin, Y. G. Guo, Adv. Energy Mater. 2018,8,
1701912.
[142] N. Hong, K. Wu, Z. Peng, Z. Zhu, G. Jia, M. Wang, J. Phys. Chem. C
2020,124, 22925.
[143] F. Leccardi, D. Nodari, D. Spada, M. Ambrosetti, M. Bini, Elec-
trochem 2021,2, 335.
[144] T. Song, L. Chen, D. Gastol, B. Dong, J. F. Marco, F. Berry, P. Slater,
D. Reed, E. Kendrick, Chem. Mater. 2022,34, 4153.
[145] L. Sun, Y. Xie, X. Z. Liao, H. Wang, G. Tan, Z. Chen, Y. Ren, J. Gim,
W. Tang, Y. S. He, K. Amine, Z. F. Ma, Small 2018,14, 1704523.
[146] C. Zhang, R. Gao, L. Zheng, Y. Hao, X. Liu, ACS Appl. Mater. Inter-
faces 2018,10, 10819.
[147] Y. Yuan, X. Wang, J. Jiang, C. Guo, D. Wang, A. Zhou, J. Electron.
Mater. 2023,52, 3509.
[148] X. Liang, H. Kim, H. G. Jung, Y.-K. Sun, Adv. Funct. Mater. 2021,31,
2008569.
[149] M. Leng, J. Bi, W. Wang, Z. Xing, W. Yan, X. Gao, J. Wang, R. Liu, J.
Alloys Compd. 2020,816, 152581.
[150] K. Kubota, N. Fujitani, Y. Yoda, K. Kuroki, Y. Tokita, S. Komaba, J.
Mater. Chem. A 2021,9, 12830.
[151] Y. Meng, J. An, L. Chen, G. Chen, L. Shi, M. Lu, D. Zhang, Chem.
Commun. 2020,56, 8079.
[152] M. Leng, C. Xia, Z. Zhou, X. Shen, J. Bi, C. Huang, Electrochim. Acta
2023,441, 141867.
[153] H. Liu, J. Xu, C. Ma, Y. S. Meng, Chem. Commun. 2015,51, 4693.
[154] J. Ren, R. Dang, Y. Yang, K. Wu, Y. Lee, Z. Hu, X. Xiao, M. Wang,
Energy Technol. 2020,8, 1901504.
[155] Q.-C. Wang, J.-K. Meng, X.-Y. Yue, Q.-Q. Qiu, Y. Song, X.-J. Wu, Z.-W.
Fu, Y.-Y. Xia, Z. Shadike, J. Wu, X.-Q. Yang, Y.-N. Zhou, J. Am. Chem.
Soc. 2019,141, 840.
[156] K. Zhang, D. Kim, Z. Hu, M. Park, G. Noh, Y. Yang, J. Zhang, V. W.-
H. Lau, S.-L. Chou, M. Cho, S.-Y. Choi, Y.-M. Kang, Nat. Commun.
2019,10, 5203.
[157] C. Zhou, L. Yang, C. Zhou, J. Liu, R. Hu, J. Liu, M. Zhu, J Energy Chem
2021,60, 341.
[158] T.-Y. Yu, Y.-K. Sun, J. Mater. Chem. A 2022,10, 23639.
[159] H.-H. Ryu, G. Han, T.-Y. Yu, Y.-K. Sun, J. Phys. Chem. C 2021,125,
6593.
[160] X. Qi, Y. Wang, L. Jiang, L. Mu, C. Zhao, L. Liu, Y.-S. Hu, L. Chen, X.
Huang, Part. Part. Syst. Charact. 2016,33, 538.
[161] Q. Wang, S. Mariyappan, J. Vergnet, A. M. Abakumov, G. Rousse,
F. Rabuel, M. Chakir, J. M. Tarascon, Adv. Energy Mater. 2019,9,
1901785.
[162] M. Sathiya, Q. Jacquet, M.-L. Doublet, O. M. Karakulina,
J. Hadermann, J.-M. Tarascon, Adv. Energy Mater. 2018,8,
1702599.
[163] H.-R. Yao, X.-G. Yuan, X.-D. Zhang, Y.-J. Guo, L. Zheng, H. Ye, Y.-X.
Yin, J. Li, Y. Chen, Y. Huang, Z. Huang, Y.-G. Guo, Energy Storage
Mater. 2023,54, 661.
[164] H.-R. Yao, P.-F. Wang, Y. Gong, J. Zhang, X. Yu, L. Gu, C. Ouyang,
Y.-X.Yin,E.Hu,X.-Q.Yang,E.Stavitski,Y.-G.Guo,L.-J.Wan,J. Am.
Chem. Soc. 2017,139, 8440.
[165] P. F. Wang, H. Xin, T. T. Zuo, Q. Li, X. Yang, Y. X. Yin, X. Gao, X. Yu,
Y. G. Guo, Angew. Chem., Int. Ed. 2018,57, 8178.
[166] P. F. Wang, H. R. Yao, X. Y. Liu, J. N. Zhang, L. Gu, X. Q. Yu, Y. X. Yin,
Y. G. Guo, Adv. Mater. 2017,29, 1700210.
[167] B. S. Kumar, A. Pradeep, V. Srihari, H. K. Poswal, R. Kumar, A.
Amardeep, A. Chatterjee, A. Mukhopadhyay, Adv. Energy Mater.
2023,13, 2204407.
[168] B. Peng, Y. Chen, L. Zhao, S. Zeng, G. Wan, F. Wang, X. Zhang, W.
Wang, G. Zhang, Energy Storage Mater. 2023,56, 631.
[169] C. Yu, L. Yang, S. Sun, D. Chen, Y. Yin, H. Y. Yang, Y. Bai, Ceram. Int.
2022,48, 36715.
[170] L. Zheng, R. Fielden, J. C. Bennett, M. N. Obrovac, J. Power Sources
2019,433, 226698.
[171] B. Mortemard De Boisse, J. H. Cheng, D. Carlier, M. Guignard, C. J.
Pan, S. Bordère, D. Filimonov, C. Drathen, E. Suard, B. J. Hwang, A.
Wattiaux, C. Delmas, J. Mater. Chem. A 2015,3, 10976.
[172] Y. Yoda, K. Kubota, K. Kuroki, S. Suzuki, K. Yamanaka, T. Yaji, S.
Amagasa, Y. Yamada, T. Ohta, S. Komaba, Small 2020,16, 2006483.
[173] P. Vassilaras, D.-H. Kwon, S. T. Dacek, T. Shi, D.-H. Seo, G. Ceder, J.
C. Kim, J. Mater. Chem. A 2017,5, 4596.
[174] S. Guo, Y. Sun, P. Liu, J. Yi, P. He, X. Zhang, Y. Zhu, R. Senga, K.
Suenaga, M. Chen, H. Zhou, Sci. Bull. 2018,63, 376.
[175] Z. Ma, Z. Zhao, H. Xu, J. Sun, X. He, Z. Lei, Z. H. Liu, R. Jiang, Q.
Li, Small 2021,17, 2006259.
[176] R. Zhang, J. Liang, C. Zeng, J. Chen, Y. Ma, T. Zhai, H. Li, Sci. China
Mater. 2023,66, 88.
[177] Z. Hu, M. Weng, Z. Chen, W. Tan, S. Li, F. Pan, Nano Energy 2021,
83, 105834.
[178] J.-Y. Hwang, C. S. Yoon, I. Belharouak, Y.-K. Sun, J. Mater. Chem. A
2016,4, 17952.
[179] M. Sathiya, K. Hemalatha, K. Ramesha, J.-M. Tarascon, A. Prakash,
Chem. Mater. 2012,24, 1846.
[180] T.-Y. Yu, J.-Y. Hwang, I. T. Bae, H.-G. Jung, Y.-K. Sun, J. Power Sources
2019,422,1.
[181] L. Tan, Z. Liu, Q. Chen, H. Yi, C. Wang, Z. Zhao, L. Song, S. Zhong,
X. Wu, L. Li, J. Colloid Interface Sci. 2022,622, 1037.
[182] D. Wu, X. Yang, S. Feng, Y. Zhu, M. Gu, Nano Lett. 2021,21, 9619.
[183] J.-Y. Hwang, S.-T. Myung, D. Aurbach, Y.-K. Sun, J. Power Sources
2016,324, 106.
[184] J. Xu, J. Chen, K. Zhang, N. Li, L. Tao, C.-P. Wong, Nano Energy 2020,
78, 105142.
[185] H.-R. Yao, W.-J. Lv, Y.-X. Yin, H. Ye, X.-W. Wu, Y. Wang, Y. Gong, Q. Li,
X. Yu, L. Gu, Z. Huang, Y.-G. Guo, ACS Appl. Mater. Interfaces 2019,
11, 22067.
[186] C. Deng, E. Gabriel, P. Skinner, S. Lee, P. Barnes, C. Ma, J. Gim, M.
L. Lau, E. Lee, H. Xiong, ACS Appl. Mater. Interfaces 2020,12, 51397.
[187] S. Xu, H. Chen, C. Li, R. Nie, Y. Yang, M. Zhou, X. Zhang, H. Zhou,
Ionics 2023,29, 1873.
[188] S.-M. Oh, S.-T. Myung, J.-Y. Hwang, B. Scrosati, K. Amine, Y.-K. Sun,
Chem. Mater. 2014,26, 6165.
[189] W. Qin, Y. Liu, J. Liu, Z. Yang, Q. Liu, Electrochim. Acta 2022,418,
140357.
[190] J. Liu, W. Qin, Z. Yang, Q. Liu, Y. Liu, J. Alloys Compd. 2022,933,
167714.
[191] T. Hwang, J. H. Lee, S. H. Choi, R. G. Oh, D. Kim, M. Cho, W. Cho,
M. S. Park, ACS Appl. Mater. Interfaces 2019,11, 30894.
[192] N. Voronina, H. J. Kim, M. Shin, S.-T. Myung, J. Power Sources 2021,
514, 230581.
Adv. Energy Mater. 2023, 2301975 © 2023 Wiley-VCH GmbH
2301975 (37 of 41)
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202301975 by Hanyang University Library, Wiley Online Library on [18/08/2023]. 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
www.advancedsciencenews.com www.advenergymat.de
[193] L. Yu, Y. X. Chang, M. Liu, Y. H. Feng, D. Si, X. Zhu, X. Z. Wang, P. F.
Wan g, S. X u, ACS Appl. Mater. Interfaces 2023,15, 23236.
[194] Y. You, S. Xin, H. Y. Asl, W. Li, P.-F. Wang, Y.-G. Guo, A. Manthiram,
Chem 2018,4, 2124.
[195] X. Gao, J. Chen, H. Liu, S. Yin, Y. Tian, X. Cao, G. Zou, H. Hou, W.
Wei, L. Chen, X. Ji, Chem. Eng. J. 2021,406, 126830.
[196] H. Fu, G. Fan, J. Zhou, X. Yu, X. Xie, J. Wang, B. Lu, S. Liang, Inorg.
Chem. 2020,59, 13792.
[197] Y. M. Zheng, X. B. Huang, X. M. Meng, S. D. Xu, L. Chen, S. B. Liu,
D. Zhang, ACS Appl. Mater. Interfaces 2021,13, 45528.
[198] Z. Zhang, Y. Liu, Z. Liu, H. Li, Y. Huang, W. Liu, D. Ruan, X. Cai, X.
Yu, J. Power Sources 2023,567, 232930.
[199] T.-T. Wei, N. Zhang, Y.-S. Zhao, Y.-R. Zhu, T.-F. Yi, Composites, Part B
2022,238, 109912.
[200] M. M. Rahman, Y. Zhang, S. Xia, W. H. Kan, M. Avdeev, L. Mu, D.
Sokaras, T. Kroll, X.-W. Du, D. Nordlund, Y. Liu, F. Lin, J. Phys. Chem.
C2019,123, 11428.
[201] C. Zhao, F. Ding, Y. Lu, L. Chen, Y. S. Hu, Angew. Chem., Int. Ed.
2020,59, 264.
[202] X. Wang, X. Li, H. Fan, M. Miao, Y. Zhang, W. Guo, Y. Fu, J Energy
Chem 2022,67, 276.
[203] J. L. Yue, W. W. Yin, M. H. Cao, S. Zulipiya, Y. N. Zhou, Z. W. Fu,
Chem. Commun. 2015,51, 15712.
[204] D. A. Anang, J.-H. Park, D. S. Bhange, M. K. Cho, W. Y. Yoon, K. Y.
Chung, K.-W. Nam, Ceram. Int. 2019,45, 23164.
[205] H. Guo, M. Avdeev, K. Sun, X. Ma, H. Wang, Y. Hu, D. Chen, Chem.
Eng. J. 2021,412, 128704.
[206] C.-C. Lin, H.-Y. Liu, J.-W. Kang, C.-C. Yang, C.-H. Li, H.-Y. T. Chen,
S.-C. Huang, C.-S. Ni, Y.-C. Chuang, B.-H. Chen, C.-K. Chang, H.-Y.
Chen, Energy Storage Mater. 2022,51, 159.
[207] F. Ding, C. Zhao, D. Xiao, X. Rong, H. Wang, Y. Li, Y. Yang, Y. Lu, Y.
S. Hu, J. Am. Chem. Soc. 2022,144, 8286.
[208] K. Walczak, A. Plewa, C. Ghica, W. Zaj ˛
ac, A. Trenczek-Zaj ˛
ac, M.
Zaj ˛
ac, J. Toboła, J. Molenda, Energy Storage Mater. 2022,47,
500.
[209] L. Yao, P. Zou, C. Wang, J. Jiang, L. Ma, S. Tan, K. A. Beyer, F. Xu, E.
Hu, H. L. Xin, Adv. Energy Mater. 2022,12, 2201989.
[210] X.-Y. Du, Y. Meng, H. Yuan, D. Xiao, Energy Storage Mater. 2023,56,
132.
[211] W. Zuo, Y. Yang, Acc Mater Res 2022,3, 709.
[212] M. Bianchini, J. Wang, R. J. Clément, B. Ouyang, P. Xiao, D. Kitchaev,
T. Shi, Y. Zhang, Y. Wang, H. Kim, M. Zhang, J. Bai, F. Wang, W. Sun,
G. Ceder, Nat. Mater. 2020,19, 1088.
[213] Q. Wang, S. Chu, S. Guo, Chin. Chem. Lett. 2020,31, 2167.
[214] J. Chen, L. Li, L. Wu, Q. Yao, H. Yang, Z. Liu, L. Xia, Z. Chen, J. Duan,
S. Zhong, J. Power Sources 2018,406, 110.
[215] B. Xiao, X. Liu, M. Song, X. Yang, F. Omenya, S. Feng, V. Sprenkle,
K. Amine, G. Xu, X. Li, D. Reed, Nano Energy 2021,89, 106371.
[216] X. Gao, H. Liu, H. Chen, Y. Mei, B. Wang, L. Fang, M. Chen, J. Chen,
J. Gao, L. Ni, L. Yang, Y. Tian, W. Deng, R. Momen, W. Wei, L. Chen,
G. Zou, H. Hou, Y.-M. Kang, X. Ji, Sci. Bull. 2022,67, 1589.
[217] G.-L. Xu, R. Amine, Y.-F. Xu, J. Liu, J. Gim, T. Ma, Y. Ren, C.-J. Sun, Y.
Liu, X. Zhang, S. M. Heald, A. Solhy, I. Saadoune, W. L. Mattis, S.-G.
Sun, Z. Chen, K. Amine, Energy Environ. Sci. 2017,10, 1677.
[218] E. Lee, J. Lu, Y. Ren, X. Luo, X. Zhang, J. Wen, D. Miller, A. Dewahl,
S. Hackney, B. Key, D. Kim, M. D. Slater, C. S. Johnson, Adv. Energy
Mater. 2014,4, 1400458.
[219] P. A. Maughan, A. B. Naden, J. T. S. Irvine, A. R. Armstrong, Com-
mun. Mater. 2023,4,6.
[220] L. Yu, Z. Cheng, K. Xu, Y.-X. Chang, Y.-H. Feng, D. Si, M. Liu, P.-F.
Wan g, S. X u, Energy Storage Mater. 2022,50, 730.
[221] Z. Cheng, X.-Y. Fan, L. Yu, W. Hua, Y.-J. Guo, Y.-H. Feng, F.-D. Ji, M.
Liu, Y.-X. Yin, X. Han, Y.-G. Guo, P.-F. Wang, Angew. Chem., Int. Ed.
2022,61, 202117728.
[222] Z.-Y. Li, J. Zhang, R. Gao, H. Zhang, L. Zheng, Z. Hu, X. Liu, J. Phys.
Chem. C 2016,120, 9007.
[223] J. Lin, Q. Huang, K. Dai, Y. Feng, X. Luo, L. Zhou, L. Chen, C. Liang,
C. Zhang, W. Wei, J. Power Sources 2022,552, 232235.
[224] S. Guo, P. Liu, H. Yu, Y. Zhu, M. Chen, M. Ishida, H. Zhou, Angew.
Chem., Int. Ed. 2015,54, 5894.
[225] C. Deng, P. Skinner, Y. Liu, M. Sun, W. Tong, C. Ma, M. L. Lau, R.
Hunt, P. Barnes, J. Xu, H. Xiong, Chem. Mater. 2018,30, 8145.
[226] D. Zhou, W. Huang, X. Lv, F. Zhao, J. Power Sources 2019,421, 147.
[227] Y. Liang, H. Xu, K. Jiang, J. Bian, S. Guo, H. Zhou, Chem. Commun.
2021,57, 2891.
[228] S.-Y. Zhang, Y.-J. Guo, Y.-N. Zhou, X.-D. Zhang, Y.-B. Niu, E.-H.
Wang, L.-B. Huang, P.-F. An, J. Zhang, X.-A. Yang, Y.-X. Yin, S. Xu,
Y. - G . G u o , Small 2021,17, 2007236.
[229] C. Zhao, Q. Wang, Z. Yao, J. Wang, B. Sánchez-Lengeling, F. Ding, X.
Qi,Y.Lu,X.Bai,B.Li,H.Li,A.Aspuru-Guzik,X.Huang,C.Delmas,
M. Wagemaker, L. Chen, Y.-S. Hu, Science 2020,370, 708.
[230] Y. Xiao, H. R. Wang, H. Y. Hu, Y. F. Zhu, S. Li, J. Y. Li, X. W. Wu, S. L.
Chou, Adv. Mater. 2022,34, 2202695.
[231] J. Zhai, H. Ji, W. Ji, R. Wang, Z. Huang, T. Yang, C. Wang, T. Zhang, Z.
Chen, W. Zhao, A. Tayal, L. Jin, J. Wang, Y. Xiao, Mater. Today Energy
2022,29, 101106.
[232] F. Ding, Q. Meng, P. Yu, H. Wang, Y. Niu, Y. Li, Y. Yang, X. Rong, X.
Liu, Y. Lu, L. Chen, Y. S. Hu, Adv. Funct. Mater. 2021,31, 2101475.
[233] R. Li, J. Gao, J. Li, H. Huang, X. Li, W. Wang, L. R. Zheng, S. M. Hao,
J. Qiu, W. Zhou, Adv. Funct. Mater. 2022,32, 2205661.
[234] A. K. Paidi, W. B. Park, P. Ramakrishnan, S. H. Lee, J. W. Lee, K. S.
Lee, H. Ahn, T. Liu, J. Gim, M. Avdeev, M. Pyo, J. I. Sohn, K. Amine,
K. S. Sohn, T. J. Shin, D. Ahn, J. Lu, Adv. Mater. 2022,34, 2202137.
[235] J. Jayachitra, J. R. Joshua, A. Balamurugan, N. Sivakumar, V.
Sharmila, S. Shanavas, M. A. Haija, M. W. Alam, A. Baqais, Ceram.
Int. 2023,49, 48.
[236] J.-Y. Hwang, T.-Y. Yu, Y.-K. Sun, J. Mater. Chem. A 2018,6, 16854.
[237] X. Li, L. Liang, M. Su, L. Wang, Y. Zhang, J. Sun, Y. Liu, L. Hou, C.
Yua n , Adv. Energy Mater. 2023,13, 2203701.
[238] H.-H. Sun, J.-Y. Hwang, C. S. Yoon, A. Heller, C. B. Mullins, ACS
Nano 2018,12, 12912.
[239] N. Li, S. Wang, E. Zhao, W. Yin, Z. Zhang, K. Wu, J. Xu, Y. Kuroiwa,
Z. Hu, F. Wang, J. Zhao, X. Xiao, J Energy Chem 2022,68, 564.
[240] H. Wang, F. Ding, Y. Wang, Z. Han, R. Dang, H. Yu, Y. Yang, Z. Chen,
Y. Li, F. Xie, S. Zhang, H. Zhang, D. Song, X. Rong, L. Zhang, J. Xu,
W. Yin, Y. Lu, R. Xiao, D. Su, L. Chen, Y.-S. Hu, ACS Energy Lett. 2023,
8, 1434.
[241] J. Yang, J.-M. Lim, M. Park, G.-H. Lee, S. Lee, M. Cho, Y.-M. Kang,
Adv. Energy Mater. 2021,11, 2102444.
[242] M. Nanthagopal, C. W. Ho, N. Shaji, G. S. Sim, M. Varun Karthik, H.
K. Kim, C. W. Lee, Nanomaterials 2022,12, 984.
[243] L. Yang, S. Sun, K. Du, H. Zhao, D. Yan, H. Y. Yang, C. Yu, Y. Bai,
Ceram. Int. 2021,47, 28521.
[244] M. Leng, J. Bi, W. Wang, Z. Xing, W. Yan, X. Gao, J. Wang, R. Liu, J.
Alloys Compd. 2021,855, 157533.
[245] J.-Y. Hwang, S.-T. Myung, J. U. Choi, C. S. Yoon, H. Yashiro, Y.-K. Sun,
J. Mater. Chem. A 2017,5, 23671.
[246] D.-Y. Hwang, S.-J. Sim, B.-S. Jin, H.-S. Kim, S.-H. Lee, ACS Appl. En-
ergy Mater. 2021,4, 1743.
[247] Y. Yu, W. Kong, Q. Li, D. Ning, G. Schuck, G. Schumacher, C. Su, X.
Liu, ACS Appl. Energy Mater. 2020,3, 933.
[248] Y. Yu, D. Ning, Q. Li, A. Franz, L. Zheng, N. Zhang, G. Ren, G.
Schumacher, X. Liu, Energy Storage Mater. 2021,38, 130.
[249] Y. You, A. Dolocan, W. Li, A. Manthiram, Nano Lett. 2019,19, 182.
[250] Y. Sun, H. Wang, D. Meng, X. Li, X. Liao, H. Che, G. Cui, F. Yu, W.
Yang, L. Li, Z.-F. Ma, ACS Appl. Energy Mater. 2021,4, 2061.
[251] N. Li, J. Ren, R. Dang, K. Wu, Y. L. Lee, Z. Hu, X. Xiao, J. Power Sources
2019,429, 38.
Adv. Energy Mater. 2023, 2301975 © 2023 Wiley-VCH GmbH
2301975 (38 of 41)
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202301975 by Hanyang University Library, Wiley Online Library on [18/08/2023]. 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
www.advancedsciencenews.com www.advenergymat.de
[252] Q. Zhang, Q.-F. Gu, Y. Li, H.-N. Fan, W.-B. Luo, H.-K. Liu, S.-X. Dou,
iScience 2019,19, 244.
[253] S. Zhao, Q. Shi, R. Qi, X. Zou, J. Wang, W. Feng, Y. Liu, X. Lu, J.
Zhang, X. Yang, Y. Zhao, Electrochim. Acta 2023,441, 141859.
[254] C. H. Jo, J. H. Jo, H. Yashiro, S. J. Kim, Y.-K. Sun, S. T. Myung, Adv.
Energy Mater. 2018,8, 1702942.
[255] Z. Zhang, L. F. Nazar, Nat. Rev. Mater. 2022,7, 389.
[256] W. Xu, R. Dang, L. Zhou, Y. Yang, T. Lin, Q. Guo, F. Xie, Z. Hu, F.
Ding, Y. Liu, Y. Liu, H. Mao, J. Hong, Z. Zuo, X. Wang, R. Yang, X.
Jin, Y. Lu, X. Rong, N. Xu, Y. S. Hu, Adv. Mater. 2023, 2301314.
[257] S. Guo, Q. Li, P. Liu, M. Chen, H. Zhou, Nat. Commun. 2017,8, 135.
[258] C. Chen, Z. Han, S. Chen, S. Qi, X. Lan, C. Zhang, L. Chen, P. Wang,
W. We i , ACS Appl. Mater. Interfaces 2020,12, 7144.
[259] X. Liang, T.-Y. Yu, H.-H. Ryu, Y.-K. Sun, Energy Storage Mater. 2022,
47, 515.
[260] M. M. Rahman, Y. Xu, H. Cheng, Q. Shi, R. Kou, L. Mu, Q. Liu, S.
Xia, X. Xiao, C.-J. Sun, D. Sokaras, D. Nordlund, J.-C. Zheng, Y. Liu,
F. Lin, Energy Environ. Sci. 2018,11, 2496.
[261] E. Gabriel, D. Hou, E. Lee, H. Xiong, Energy Sci. Eng. 2022,10, 1672.
[262] H. Xu, Q. Yan, W. Yao, C.-S. Lee, Y. Tang, Small Struct. 2022,3,
2100217.
[263] L. Mu, M. M. Rahman, Y. Zhang, X. Feng, X.-W. Du, D. Nordlund, F.
Lin, J. Mater. Chem. A 2018,6, 2758.
[264] U.-H. Kim, S.-B. Lee, N.-Y. Park, S. J. Kim, C. S. Yoon, Y.-K. Sun, ACS
Energy Lett. 2022,11, 3880.
[265] J.-Y. Hwang, S.-M. Oh, S.-T. Myung, K. Y. Chung, I. Belharouak, Y.-K.
Sun, Nat. Commun. 2015,6, 6865.
[266] J.-Y. Hwang, S.-T. Myung, C. S. Yoon, S.-S. Kim, D. Aurbach, Y.-K.
Sun, Adv. Funct. Mater. 2016,26, 8083.
[267] Y. Chen, S. Wang, Y. Jie, Z. Lei, R. Cao, S. Jiao, Chem. Res. Chin. Univ.
2021,37, 280.
[268] J. Deng, W.-B. Luo, X. Lu, Q. Yao, Z. Wang, H.-K. Liu, H. Zhou, S.-X.
Dou, Adv. Energy Mater. 2018,8, 1701610.
[269] Q. Mao, R. Gao, Q. Li, D. Ning, D. Zhou, G. Schuck, G. Schumacher,
Y. H a o , X . L i u , Chem. Eng. J. 2020,382, 122978.
[270] T. Yuan, S. Li, Y. Sun, J. H. Wang, A. J. Chen, Q. Zheng, Y. Zhang,
L. Chen, G. Nam, H. Che, J. Yang, S. Zheng, Z. F. Ma, M. Liu, ACS
Nano 2022,16, 18058.
[271] J. Wang, X. He, D. Zhou, F. Schappacher, X. Zhang, H. Liu, M. C.
Stan, X. Cao, R. Kloepsch, M. S. Sofy, G. Schumacher, J. Li, J. Mater.
Chem. A 2016,4, 3431.
[272] M. Palluzzi, L. Silvestri, A. Celeste, M. Tuccillo, A. Latini, S. Brutti,
Crystal s 2022,12, 885.
[273] V. K. Kumar, S. Ghosh, S. Biswas, S. K. Martha, J. Electrochem. Soc.
2020,167, 080531.
[274] Y. Zhang, J. Wang, L. Wang, L. Duan, G. Zhang, F. Zhao, X. Zhang,
W. Lü, J. Mater. Sci. 2020,55, 13102.
[275] L. Wu, S. Guo, X. Pu, H. Yue, H. Li, P. Li, W. Li, K. Cai, W. Ding, L. Li,
Y. Zhang, W. Fa, C. Yang, Z. Zheng, W. He, Y. Cao, ACS Appl. Energy
Mater. 2022,5, 4505.
[276] D. Meghnani, R. K. Singh, Electrochim. Acta 2022,419, 140403.
[277] M. W. Alam, A. Baqais, I. Nahvi, A. Yasin, T. A. Mir, S. Shajahan,
Inorganics 2023,11, 37.
[278] Y. Xiao, P. F. Wang, Y. X. Yin, Y. F. Zhu, Y. B. Niu, X. D. Zhang, J.
Zhang,X.Yu,X.D.Guo,B.H.Zhong,Y.G.Guo,Adv. Mater. 2018,
30, 1803765.
[279] X. Luo, Q. Huang, Y. Feng, L. Zhou, W. Wei, ChemElectroChem 2022,
9, 202200821.
[280] Y. Chang, Y. Zhou, Z. Wang, X. Li, D. Wang, J. Duan, J. Wang, G. Yan,
J. Alloys Compd. 2022,922, 166283.
[281] Y. Li, X. Liang, G. Zhong, C. Wang, S. Wu, K.-Q. Xu, C. Yang, ACS
Appl. Mater. Interfaces 2020,12, 25920.
[282] G.-M. Han, Y.-S. Kim, H.-H. Ryu, Y.-K. Sun, C. S. Yoon, ACS Energy
Lett. 2022,7, 2919.
[283] Y. Zhuang, J. Zhao, Y. Zhao, X. Zhu, H. Xia, Sustainable Mater. Tech-
nol. 2021,28, e00258.
[284] J. Lamb, K. Jarvis, A. Manthiram, Small 2022,18, 2106927.
[285] J. Darga, A. Manthiram, ACS Appl. Mater. Interfaces 2022,14, 52729.
[286] V. Pamidi, S. Trivedi, S. Behara, M. Fichtner, M. A. Reddy, iScience
2022,25, 104205.
[287] L. Cao, X. Liang, X. Ou, X. Yang, Y. Li, C. Yang, Z. Lin, M. Liu, Adv.
Funct. Mater. 2020,30, 1910732.
[288] H. Zhang, L. Wang, H. Li, X. He, ACS Energy Lett. 2021,6, 3719.
[289] K. Subramanyan, V. Aravindan, ACS Energy Lett. 2023,8, 436.
[290] Y. Wan, Y. Liu, D. Chao, W. Li, D. Zhao, Nano Mater. Sci. 2022,5, 189.
[291] S. Huang, Y. Lv, W. Wen, T. Xue, P. Jia, J. Wang, J. Zhang, Y. Zhao,
Mater. Today Energy 2021,20, 100673.
[292] Q. Su, S. Zeng, M. Tang, Z. Zheng, Z. Wang, Energy Technol. 2021,
9, 2100049.
[293] U. Ghani, N. Iqbal, A. A. Aboalhassan, C. Zhou, B. Liu, J. Li, Y. Fang,
T. Aftab, J. Gu, Q. Liu, ACS Appl. Mater. Interfaces 2022,14, 47507.
[294] X. Chen, Y. Fang, H. Lu, H. Li, X. Feng, W. Chen, X. Ai, H. Yang, Y.
Cao, Small 2021,17, 2102248.
[295] L. Xiao, H. Lu, Y. Fang, M. L. Sushko, Y. Cao, X. Ai, H. Yang, J. Liu,
Adv. Energy Mater. 2018,8, 1703238.
[296] A. Kamiyama, K. Kubota, D. Igarashi, Y. Youn, Y. Tateyama, H. Ando,
K. Gotoh, S. Komaba, Angew. Chem., Int. Ed. 2021,60, 5114.
[297] Y. Li, Y. Lu, Q. Meng, A. C. S. Jensen, Q. Zhang, Q. Zhang, Y. Tong,
Y. Qi, L. Gu, M. M. Titirici, Y.-S. Hu, Adv. Energy Mater. 2019,9,
1902852.
[298] L. Yang, M. Hu, Q. Lv, H. Zhang, W. Yang, R. Lv, Carbon 2020,163,
288.
[299] Y. Lu, C. Zhao, X. Qi, Y. Qi, H. Li, X. Huang, L. Chen, Y.-S. Hu, Adv.
Energy Mater. 2018,8, 1800108.
[300] Y. Li, Y. Yuan, Y. Bai, Y. Liu, Z. Wang, L. Li, F. Wu, K. Amine, C. Wu, J.
Lu, Adv. Energy Mater. 2018,8, 1702781.
[301] F. Sun, H. Wang, Z. Qu, K. Wang, L. Wang, J. Gao, J. Gao, S. Liu, Y.
Lu, Adv. Energy Mater. 2021,11, 2002981.
[302] A. Beda, F. Rabuel, M. Morcrette, S. Knopf, P.-L. Taberna, P. Simon,
C. Matei Ghimbeu, J. Mater. Chem. A 2021,9, 1743.
[303] J. Zhao, X. X. He, W. H. Lai, Z. Yang, X. H. Liu, L. Li, Y. Qiao, Y. Xiao,
L. Li, X. Wu, S. L. Chou, Adv. Energy Mater. 2023,13, 2300444.
[304] X. X. He, J. H. Zhao, W. H. Lai, R. Li, Z. Yang, C. M. Xu, Y. Dai, Y.
Gao, X. H. Liu, L. Li, G. Xu, Y. Qiao, S. L. Chou, M. Wu, ACS Appl.
Mater. Interfaces 2021,13, 44358.
[305] Q. Gan, N. Qin, S. Gu, Z. Wang, Z. Li, K. Liao, K. Zhang, L. Lu, Z.
Xu, Z. Lu, Small Methods 2021,5, 2100580.
[306] W. Li, X. Guo, K. Song, J. Chen, J. Zhang, G. Tang, C. Liu, W. Chen,
C. Shen, Adv. Energy Mater. 2023,13, 2300648.
[307] M. Zhang, Y. Li, F. Wu, Y. Bai, C. Wu, Nano Energy 2021,82,
105738.
[308] Y. Liao, F. Luo, T. Lyu, M. Chen, C. Liu, D. Xu, P. Chen, Q. Liu, Z.
Wang, S. Li, Y. Ye, D. Wang, C. Miao, Z. Liu, D. Wang, Z. Zheng,
Diamond Relat. Mater. 2022,129, 109392.
[309] Y. Li, L. Mu, Y.-S. Hu, H. Li, L. Chen, X. Huang, Energy Storage Mater.
2016,2, 139.
[310] X. Li, J. Sun, W. Zhao, Y. Lai, X. Yu, Y. Liu, Adv. Funct. Mater. 2022,
32, 2106980.
[311] Z. Lu, J. Wang, W. Feng, X. Yin, X. Feng, S. Zhao, C. Li, R. Wang, Q.
A. Huang, Y. Zhao, Adv. Mater. 2023,35, 2211461.
[312] C. Yu, Y. Li, H. Ren, J. Qian, S. Wang, X. Feng, M. Liu, Y. Bai, C. Wu,
Carbon Energy 2023,5, e220.
[313] Y. Zhang, N. Zhang, W. Chen, Z. Rao, J. Wu, L. Xue, W. Zhang, Energy
Technol. 2019,7, 1900779.
[314] Z. Tian, Y. Zou, G. Liu, Y. Wang, J. Yin, J. Ming, H. N. Alshareef, Adv.
Sci. 2022,9, 2201207.
[315] A. Ponrouch, E. Marchante, M. Courty,J.-M. Tarascon, M. R. Palacín,
Energy Environ. Sci. 2012,5, 8572.
Adv. Energy Mater. 2023, 2301975 © 2023 Wiley-VCH GmbH
2301975 (39 of 41)
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202301975 by Hanyang University Library, Wiley Online Library on [18/08/2023]. 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
www.advancedsciencenews.com www.advenergymat.de
[316] G. G. Eshetu, S. Grugeon, H. Kim, S. Jeong, L. Wu, G. Gachot,
S. Laruelle, M. Armand, S. Passerini, ChemSusChem 2016,9,
462.
[317] Y. Huang, L. Zhao, L. Li, M. Xie, F. Wu, R. Chen, Adv. Mater. 2019,
31, 1808393.
[318] G. Kamath, R. W. Cutler, S. A. Deshmukh, M. Shakourian-
Fard, R. Parrish, J. Huether, D. P. Butt, H. Xiong, S. K. R. S.
Sankaranarayanan, J. Phys. Chem. C 2014,118, 13406.
[319] Nagmani, A. Kumar, S. Puravankara, Battery Energy 2022,1,
20220007.
[320] D. Pahari, A. Chowdhury, D. Das, T. Paul, S. Puravankara, J. Solid
State Electrochem. 2023,27, 2067.
[321] Y. Lee, J. Lee, H. Kim, K. Kang, N.-S. Choi, J. Power Sources 2016,320,
49.
[322] M. Dahbi, T. Nakano, N. Yabuuchi, S. Fujimura, K. Chihara, K.
Kubota, J.-Y. Son, Y.-T. Cui, H. Oji, S. Komaba, ChemElectroChem
2016,3, 1856.
[323] S. Wu, B. Su, K. Ni, F. Pan, C. Wang, K. Zhang, D. Y. W. Yu, Y. Zhu,
W. Zha n g , Adv. Energy Mater. 2021,11, 2002737.
[324] J. Feng, Y. An, L. Ci, S. Xiong, J. Mater. Chem. A 2015,3, 14539.
[325] H. He, D. Sun, Y. Tang, H. Wang, M. Shao, Energy Storage Mater.
2019,23, 233.
[326] J. J. Fan, P. Dai, C. G. Shi, Y. Wen, C. X. Luo, J. Yang, C. Song, L.
Huang, S. G. Sun, Adv. Funct. Mater. 2021,31, 2010500.
[327] R. Mogensen, D. Brandell, R. Younesi, ACS Energy Lett. 2016,1,
1173.
[328] D. Wu, C. Zhu, M. Wu, H. Wang, J. Huang, D. Tang, J. Ma, Angew.
Chem., Int. Ed. 2022,61, e202214198.
[329] H. Li, H. Chen, X. Shen, X. Liu, Y. Fang, F. Zhong, X. Ai, H. Yang, Y.
Cao, ACS Appl. Mater. Interfaces 2022,14, 43387.
[330] Z. Lu, H. Yang, Q. H. Yang, P. He, H. Zhou, Angew. Chem., Int. Ed.
2022,61, e202200410.
[331] K. Du, C. Wang, P. Balaya, S. R. Gajjela, M. Law, Chem. Commun.
2022,58, 533.
[332] X. Liu, X. Zheng, Y. Deng, X. Qin, Y. Huang, Y. Dai, W. Wu, Z. Wang,
W. Luo, Adv. Funct. Mater. 2022,32, 2109378.
[333] Y. Sun, P. Shi, H. Xiang, X. Liang, Y. Yu, Small 2019,15, 1805479.
[334] Y. Jin, Y. Xu, P. M. L. Le, T. D. Vo, Q. Zhou, X. Qi, M. H. Engelhard,
B. E. Matthews, H. Jia, Z. Nie, C. Niu, C. Wang, Y. Hu, H. Pan, J.-G.
Zhang, ACS Energy Lett. 2020,5, 3212.
[335] Z. Yang, J. He, W. H. Lai, J. Peng, X. H. Liu, X. X. He, X. F. Guo, L. Li,
Y. Qiao, J. M. Ma, M. Wu, S. L. Chou, Angew. Chem., Int. Ed. 2021,
60, 27086.
[336] K. Sirengo, A. Babu, B. Brennan, S. C. Pillai, J Energy Chem 2023,81,
321.
[337] H. Dai, Y. Chen, W. Xu, Z. Hu, J. Gu, X. Wei, F. Xie, W. Zhang, W. Wei,
R. Guo, G. Zhang, Energy Technol. 2020,9, 2000682.
[338] M. J. Piernas-Muñoz, E. Castillo-Martínez, J. L. Gómez-Cámer, T.
Rojo, Electrochim. Acta 2016,200, 123.
[339] J. Park, K. Ku, S.-B. Son, J. Gim, Y. Kim, E. Lee, C. Johnson, J. Elec-
trochem. Soc. 2022,169, 030536.
[340] G. Yan, D. Alves-Dalla-Corte, W. Yin, N. Madern, G. Gachot, J.-M.
Tarascon, J. Electrochem. Soc. 2018,165, A1222.
[341] G. Hernández, R. Mogensen, R. Younesi, J. Mindemark, Batteries
Supercaps 2022,5, 202100373.
[342] S. Zhang, X. Li, Y. Su, Y. Yang, H. Yu, H. Wang, Q. Zhang, Y. Lu, D.
Song, S. Wang, Q. Zhang, X. Rong, L. Zhang, L. Chen, Y. S. Hu, Adv.
Funct. Mater. 2023, 2301568.
[343] H.-J. Noh, S. Youn, C. S. Yoon, Y.-K. Sun, J. Power Sources 2013,233,
121.
Xinghui Liang received his Ph.D. degree (2023) in energy engineering from Hanyang University.He is
currently a postdoctoral researcher in the group of Professor Yang-Kook Sun at Hanyang University.
His research interests include the preparation and characterization of advanced cathode materials for
sodium-ion batteries.
Jang-Yeon Hwang received his Ph.D. degree (2018) at Hanyang University in Seoul. He is an associate
professor in the Department of Energy Engineering & Battery Engineering at Hanyang University in
Seoul. His research is mainly focused on the synthesis and characterization of advanced electrode
materials for the high-energy alkali-ion and alkali-metal batteries.
Adv. Energy Mater. 2023, 2301975 © 2023 Wiley-VCH GmbH
2301975 (40 of 41)
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202301975 by Hanyang University Library, Wiley Online Library on [18/08/2023]. 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
www.advancedsciencenews.com www.advenergymat.de
Yang-Kook Sun is currently a distinguished professor at Hanyang University in Seoul. After receiving
his Ph.D. at Seoul National University in 1992, he has pursued his passion for energy storage and
conversion materials for innovative and high-level scientific work. One of his major achievements is
the discovery of a new concept of layered concentration gradient NCM cathode materials for lithium-
ion batteries. His major research interests are design, synthesis of novel materials for lithium-ion
batteries, sodium-ion batteries, lithium-sulfur batteries, and all-solid-state batteries.
Adv. Energy Mater. 2023, 2301975 © 2023 Wiley-VCH GmbH
2301975 (41 of 41)
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202301975 by Hanyang University Library, Wiley Online Library on [18/08/2023]. 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
... There is an urgent need to find new alternatives of LIBs. The SIBs and PIBs have been largely developed in recent years due to their lower production cost, abundant resources, and similar physicochemical properties to lithium [11][12][13][14][15]. Unfortunately, the large ionic radius (Na + 1.02 Å, K + 1.38 Å) severely hinder the Na + and K + diffusion kinetics behavior, which negatively affects the cycling performance and rate performance of the device [16][17][18]. During the charging/discharging process, the repeated embedding/de-embedding of Na + /K + will very easily result in the fragmentation of the anode material, which will lead to the deterioration of the electrochemical reversibility [19,20]. ...
Article
Full-text available
Sodium-ion batteries (SIBs) and potassium-ion batteries (PIBs) are potential alternatives of lithium-ion batteries (LIBs) due to the abundant and cost-effective availability of sodium and potassium resources. Unfortunately, they are difficult to use for large-scale grid energy storage due to the lack of suitable anode materials for sodium/potassium energy storage. Biomass-derived carbon, which is widely available and environmentally friendly, is one of the most promising anode materials for SIBs/PIBs, but the design and regulation of its microstructure is exceptionally complex. By selecting suitable biomass precursors, it is expected that biomass-derived carbon with suitable microstructures can be simply prepared. In this study, Wedelia chinensis were selected as biomass precursors, and biomass-derived carbon materials with large interfacial spacing, suitable pores and high-specific surface area were prepared by a simple one-step pyrolysis method. The material exhibited fast energy storage kinetics when electrochemically tested as an anode and showed different performance advantages in storing sodium/potassium. When tested as an anode for SIBs, it exhibited excellent specific capacity and cycling stability (380.7 mA h g⁻¹ after 500 cycles at 100 mA g⁻¹); When tested as an anode for PIBs, it exhibited excellent rate performance (128.6 mA h g⁻¹ at 10 A g⁻¹).
... V vs. SHE (standard hydrogen electrode) compared to -3.01 V vs. SHE for Li + /Li], SIBs commonly show lower specific energy. Also, from the perspective of intercalation reaction, the much larger size of Na + (1.02 Å vs. 0.76 Å for Li + ) tends to cause sluggish solid-state diffusion and larger dimensional changes of the host structures [9,10] . Therefore, the development of SIBs with high specific capacity, high operating voltage, high rate capability, and long lifespan remains challenging. ...
Article
Full-text available
Sodium-ion batteries (SIBs) have attracted enormous attention as candidates in stationary energy storage systems, because of the decent electrochemical performance based on cheap and abundant Na-ion intercalation chemistry. Layered oxides, the workhorses of modern lithium-ion batteries, have regained interest for replicating their success in enabling SIBs. A unique feature of sodium layered oxides is their ability to crystallize into a thermodynamically stable P2-type layered structure with under-stoichiometric Na content. This structure provides highly open trigonal prismatic environments for Na ions, permitting high Na+ mobility and excellent structural stability. This review delves into the intrinsic characteristics and key challenges faced by P2-type cathodes and then comprehensively summarizes the up-to-date advances in modification strategies from compositional design, elemental doping, phase mixing, morphological control, and surface modification to sodium compensation. The updated understanding presented in this review is anticipated to guide and expedite the development of P2-type layered oxide cathodes for practical SIB applications.
Article
Sodium-ion batteries (SIBs), which serve as alternatives or supplements to lithium-ion batteries, have been developed rapidly in recent years. Designing advanced high-performance layered NaxTMO2 cathode materials is beneficial for accelerating the commercialization of SIBs. Herein, the recent research progress on scalable synthesis methods, challenges on the path to commercialization and practical material design strategies for layered NaxTMO2 cathode materials is summarized. Co-precipitation method and solid-phase method are commonly used to synthesize NaxTMO2 on mass production and show their own advantages and disadvantages in terms of manufacturing cost, operative difficulty, sample quality and so on. To overcome drawbacks of layered NaxTMO2 cathode materials and meet the requirements for practical application, a detailed and deep understanding of development trends of layered NaxTMO2 cathode materials is also provided, including high specific energy materials, high-entropy oxides, single crystal materials, wide operation temperature materials and high air stability materials. This work can provide useful guidance in developing practical layered NaxTMO2 cathode materials for commercial SIBs.
Article
O3-type layered oxide cathodes are promising in practical sodium-ion batteries (SIBs) owing to their high theoretical capacity, facile synthesis, and sufficient Na+ storage. However, they face challenges such as rapid...
Article
Full-text available
The successful utilization of innovative sulfide‐based solid‐state batteries in energy storage hinges on developing scalable technologies and machinery for upscaling their production. While multiple Gigafactories for lithium‐ion batteries are already operational worldwide, the upscaling of solid‐state batteries exhibiting their full potential remains to be seen in the near future. In this study, the conventional production of lithium‐ion batteries is reconsidered, and the feasibility of seamlessly integrating sulfide‐based solid‐state batteries into the existing process chains is discussed. Scalable technologies and key challenges along the process chain of sulfide‐based solid‐state batteries are accordingly addressed. Experimental investigations yield crucial insights into enabling large‐scale production of sulfide‐based battery components while highlighting remaining challenges from a production perspective. An overview of the roll‐to‐roll machinery housed in microenvironments under an inert atmosphere in the “Sulfidic Cell Production Advancement Center” at the Institute for Machine Tools and Industrial Management at the Technical University of Munich is given.
Article
Full-text available
The development of large‐scale energy storage systems (ESSs) aimed at application in renewable electricity sources and in smart grids is expected to address energy shortage and environmental issues. Sodium‐ion batteries (SIBs) exhibit remarkable potential for large‐scale ESSs because of the high richness and accessibility of sodium reserves. Using low‐cost and abundant elements in cathodes with long cycling stability is preferable for lowering expenses on cathodes. Many investigated cathodes for SIBs are dogged by structural and morphology changes, unstable interphases between the cathode and the electrolyte, and air sensitivity, causing unsatisfactory cycling performance. Therefore, understanding the mechanism of capacity degeneration in depth and developing precise solutions are critical for designing low‐cost cathodes that are highly stable under cycling. Herein, recent progress in long‐cycle‐life and low‐cost cathodes for SIBs is focused on, and a comprehensive discussion of the key points in SIBs toward large‐scale applications is provided. The roots of the unstable cycling performance of low‐cost cathodes are discussed. Also, effective strategies are summarized from the recent progress on long‐cycle‐life and low‐cost cathodes. This review is expected to encourage deeper investigation of long‐lifespan cathodes for SIBs, particularly for potential large‐scale industrialization.
Article
Full-text available
High‐performance layered oxide cathode materials are vital to the industrialization of Na‐ion batteries (NIBs) due to their high theoretical capacity and facile production. These materials, however, suffer from complicated phase transitions and poor capacity retention cycling. The modifying impacts of classic strategies, such as doping and coating, are generally restricted. Herein, a rational four‐in‐one strategy to boost the overall performance of Nax[Ni,Mn]O2 materials by innovative composition design and elegant synthesis management is provided. After modification, the half‐cell capacity retention rate increases from 45% to 77% after 250 cycles at 1 C rate. The capacity of the full cells is maintained at 83% after 300 cycles at 0.5 C rate. This four‐in‐one strategy, which encompasses optimization of the energy band structure, ion diffusion, crystal structure, and particle morphology, sheds new insight into the overall optimization of cathode materials for NIBs.
Article
Full-text available
Hard carbon is one of the most promosing anodes for resource‐rich sodium‐ion batteries. However, an unsatisfactory solid–electrolyte‐interphase formed by irreversible electrolyte consumption caused by defects or oxygen‐containing functional groups of hard carbon impedes its further application. Herein, a novel composite binder that is composed of polar polymer chondroitin sulfate A (sodium salt) and polyethylene oxide by hydrogen bonding demonstrates defect passivation capability. This composite binder can reduce the exposure of defects by forming hydrogen bonds with oxygen functional groups on the hard carbon surface and inhibit the decomposition of electrolyte confirmed by in situ differential electrochemical mass spectrometry. In situ Raman and theoretical calculations reveal that multiple polar functional groups in chondroitin sulfate A (sodium salt) can accelerate the transport of Na⁺ by adsorbing and facilitate the decomposition of PF6⁻ to form NaF. Additionally, polyethylene oxide in the composite binder can increase viscosity and accelerate the transport of Na⁺. As a result, an ultra‐thin (9 nm, cyro‐TEM) and NaF‐rich solid–electrolyte interphase is obtained, thereby the hard carbon anode achieves improved initial Coulombic efficiency (84%) and high‐capacity retention of 94% after 150 cycles in a NaPF6/ethylene carbonate/dimethyl carbonate electrolyte.
Article
Full-text available
The deposition of volatilized Na ⁺ on the surface of the cathode during sintering results in the formation of surface residual alkali (NaOH/Na 2 CO 3 /NaHCO 3 ) in layered cathode materials, leading to serious interfacial reactions and performance degradation. This phenomenon is particularly evident in O3‐NaNi 0.4 Cu 0.1 Mn 0.4 Ti 0.1 O 2 (NCMT). In this study, we propose a strategy to transform waste into treasure by converting residual alkali into a solid electrolyte. Mg(CH 3 COO) 2 and H 3 PO 4 are reacted with surface residual alkali to generate the solid electrolyte NaMgPO 4 on the surface of NCMT, which can be labeled as NaMgPO 4 @NaNi 0.4 Cu 0.1 Mn 0.4 Ti 0.1 O 2 ‐X (NMP@NCMT‐X, where X indicates the different amounts of Mg ²⁺ and PO 4 ³⁻ ). NaMgPO 4 acts as a special ionic conductivity channel on the surface to improve the kinetics of the electrode reactions, remarkably improving the rate capability of the modified cathode at a high current density in the half‐cell. Additionally, NMP@NCMT‐2 enables a reversible phase transition from the P3 to OP2 phase in the charge–discharge process above 4.2 V and achieves a high specific capacity of 157.3 mAh g ⁻¹ and outstanding capacity retention in the full cell. The strategy can effectively and reliably stabilize the interface and improve the performance of layered cathodes for Na‐ion batteries (NIBs). This article is protected by copyright. All rights reserved
Article
Full-text available
Hard carbon (HC) anodes have shown extraordinary promise for sodium‐ion batteries, but are limited to their poor initial coulombic efficiency (ICE) and low practical specific capacity due to the large amount of defects. These defects with oxygen containing groups cause irreversible sites for Na⁺ ions. Highly graphited carbon decreases defects, while potentially blocking diffusion paths of Na⁺ ions. Therefore, molecular‐level control of graphitization of hard carbon with open accessible channels for Na⁺ ions is key to achieve high‐performance hard carbon. Moreover, it is challenging to design a conventional method to obtain HCs with both high ICE and capacity. Herein, a universal strategy is developed as manganese ions‐assisted catalytic carbonization to precisely tune graphitization degree, eliminate defects, and maintain effective Na⁺ ions paths. The as‐prepared hard carbon has a high ICE of 92.05% and excellent cycling performance. Simultaneously, a sodium storage mechanism of “adsorption‐intercalation‐pore filling‐sodium cluster formation” is proposed, and a clear description given of the boundaries of the pore structure and the specific dynamic process of pore filling.
Article
Full-text available
This work evolves a universal strategy, toward simultaneously addressing the air/water‐instability and structural‐cum‐electrochemical instability of “layered” Na–transition‐metal (TM)–oxide‐based cathode materials for Na‐ion batteries, by way of varying the “interslab” spacing via tuning the TMO bond covalency. In this regard, model O3‐structured NaTMoxides, with varied “charge‐to‐size” ratio of the cation‐combination (viz., TM‐ + non‐TM‐ions) in the TM‐layer [i.e., (C:S)TM], are designed and subjected to structural characterizations, density‐functional‐theory‐based studies, air/water‐exposure studies, electrochemical cycling, and operando investigations. Such studies have yielded a clear correlation‐cum‐trend concerning lower (C:S)TM => lower TMO covalency => higher effective negative charge on O‐ion (which gets shared by both TM‐ and Na‐ions) => stronger‐cum‐shorter NaO bond => reduced “interslab” spacing => lower Na‐transport kinetics => suppressed spontaneous Na‐extraction upon air/water‐exposure => concomitant vastly improved air/water‐stability => suppressed/delayed O3 → P3 transformation during electrochemical Na‐extraction (i.e., charging) => concomitant vastly improved electrochemical cyclic stability. Furthermore, a critical d(ONaO)/d(OTMO) of ≈1.38 for the O3 structure, corresponding to the initiation of O3 → P3 transformation during desodiation/charging is revealed. NaTMO2s having higher initial (C:S)TMs reach this critical d(ONaO)/d(OTMO) earlier (i.e., upon minimized Na‐removal) and, thus, suffer from more transformations during continued desodiation/charge, resulting in structural‐cum‐electrochemical instability.
Article
Full-text available
The introduction of copper (Cu) element to iron-manganese-based layered cathode materials can effectively enhance their cycling stability and air tolerance. However, the low redox reactivity of Cu2+ decreases the capacity of the copper-iron-manganese layered oxide cathode material. Recently, lithium (Li) doping has been regarded as an efficient strategy to exploit high-capacity cathode materials by enabling high-covalency transition metals. Here, we report a Na1.0LixCu0.22Fe0.30Mn0.48O2 (x = 0.025, 0.05, 0.075) cathode material with increased capacity by adding Li into a Na1.0Cu0.22Fe0.30Mn0.48O2 cathode via a simple solid-phase sintering method. The doped Li element can regulate the redox reactivities of the adjacent Fe and Mn elements, leading to the promotion of the Fe redox reactivity and the suppression of Mn redox reactivity, which prevents both the Jahn–Teller effect and the structure collapse during the charge/discharge process. In conclusion, Li doping can not only improve the capacity of the cathode material but also improve its stability. When x = 0.075, the capacity of Na1Li0.075Cu0.22Fe0.30Mn0.48O2 cathode can reach 114.2 mAh g−1 with a high capacity retention of 90.2% after 300 cycles at 1 C. These results shed light on the role play of Li in the transition metal layer, and can guide the design and modification for high-performance SIBs of layered materials.
Article
O3-type layered oxides with high initial sodium content are promising cathode candidates for Na-ion batteries. However, affected by the undesired transition metal slab sliding and reaction with H2O/CO2, their further application is typically hindered by unsatisfactory cycling stability upon charging to high voltage and poor storage stability under humid air. Herein, we demonstrate a Fe/Ti cosubstitution strategy to simultaneously enhance the electrochemical performance and storage stability of pristine O3-NaNi0.5Mn0.5O2 cathode material, via employing high redox potential and inactive stabilized dopants. The resultant Fe/Ti cosubstituted Na0.95Ni0.40Fe0.15Mn0.3Ti0.15O2 undergoes highly reversible O3-P3-OP2 phase transitions with a small cell volume change of 2.8%, instead of complex O3-O'3-P3-P'3-P3'-O1 phase transitions in NaNi0.5Mn0.5O2. Consequently, the cathode displays a high specific capacity of 161.6 mAh g-1 with an average working voltage of 3.28 V and 81.8% capacity retention after 200 cycles at 5C. Furthermore, the cathode material remains very stable after exposure to air for 7 days and even after soaking in water for 1 h, owing to the prohibition of sodium losing by elevating redox potential and contracting sodium layer spacing. This work proposes an effective method to enhance the electrochemical performance and storage stability of O3-type layered oxide cathodes and promises advancing Na-ion batteries toward large-scale industrialization.