ArticlePDF AvailableLiterature Review

Electrochemical coupling in subnanometer pores/channels for rechargeable batteries

Authors:

Abstract

This review categorizes subnanometer pores/channels (SNPCs) from structural perspective and demonstrates electrochemical couplings in SNPCs for batteries while proposing corresponding challenges and future research directions.
This journal is © The Royal Society of Chemistry 2024 Chem. Soc. Rev.
Cite this: DOI: 10.1039/d3cs01043k
Electrochemical coupling in subnanometer
pores/channels for rechargeable batteries
Yao-Jie Lei,
a
Lingfei Zhao,
b
Wei-Hong Lai,
b
Zefu Huang,
a
Bing Sun,
a
Pauline Jaumaux,
a
Kening Sun,*
c
Yun-Xiao Wang*
d
and Guoxiu Wang *
a
Subnanometer pores/channels (SNPCs) play crucial roles in regulating electrochemical redox reactions
for rechargeable batteries. The delicately designed and tailored porous structure of SNPCs not only
provides ample space for ion storage but also facilitates efficient ion diffusion within the electrodes in
batteries, which can greatly improve the electrochemical performance. However, due to current
technological limitations, it is challenging to synthesize and control the quality, storage, and transport of
nanopores at the subnanometer scale, as well as to understand the relationship between SNPCs and
performances. In this review, we systematically classify and summarize materials with SNPCs from a
structural perspective, dividing them into one-dimensional (1D) SNPCs, two-dimensional (2D) SNPCs,
and three-dimensional (3D) SNPCs. We also unveil the unique physicochemical properties of SNPCs and
analyse electrochemical couplings in SNPCs for rechargeable batteries, including cathodes, anodes,
electrolytes, and functional materials. Finally, we discuss the challenges that SNPCs may face in
electrochemical reactions in batteries and propose future research directions.
1. Introduction
The endeavours to achieve high energy density, high power
density, and high cycling stability for various rechargeable
battery systems have been the persistent pursuit of the research
community for decades.
1–4
Among the diverse approaches,
reducing the size of electrode materials with essential phase
engineering for cathodes and morphology design for anodes is
one of the most effective strategies to enhance the electroche-
mical performance of the batteries.
5,6
With the advancement of
systematic characterization techniques for optimal electrode
a
Centre for Clean Energy Technology, School of Mathematical and Physical
Sciences, Faculty of Science, University of Technology Sydney, Sydney, NSW 2007,
Australia. E-mail: Guoxiu.Wang@uts.edu.au
b
Institute for Superconducting & Electronic Materials, Australian Institute of
Innovative Materials, University of Wollongong, Innovation Campus, Squires Way,
North Wollongong, NSW 2500, Australia
c
School of Chemistry and Chemical Engineering, Beijing Institute of Technology,
Beijing 10081, P. R. China. E-mail: sunkn@bit.edu.cn
d
Institute of Energy Materials Science (IEMS), University of Shanghai for Science
and Technology, 516 Jungong Road, Shanghai, 200093, P. R. China.
E-mail: yunxiaowang@usst.edu.cn
Yao-Jie Lei
Yao-Jie Lei is a Research Associate
working under the supervision of
Prof. Guoxiu Wang at the Centre
for Clean Energy Technology,
University of Technology, Sydney.
He received his bachelor of science
major in chemistry from Lanzhou
University, and research master
degree from the University of
Sydney. He obtained his PhD from
the University of Wollongong. His
research interests include the
synthesis of nanostructured mater-
ials for room-temperature Na–S
batteries.
Lingfei Zhao
Lingfei Zhao is currently an
associate research fellow at the
Institute for Superconducting &
Electronic Materials, University
of Wollongong, where he earned
his PhD in July 2023. His
research interest focuses on
materials design and electrolyte
engineering for high energy
density rechargeable batteries,
including Li/Na-ion batteries
and Li/Na–sulfur batteries.
These authors contributed equally to this work.
Received 27th November 2023
DOI: 10.1039/d3cs01043k
rsc.li/chem-soc-rev
Chem Soc Rev
REVIEW ARTICLE
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
View Journal
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2024
materials and in-depth understanding of the electrochemical
mechanisms, the fact that the electrode materials with fine
structures, especially those exhibiting variations at the subnan-
ometer scale, lead to substantial enhancement in electroche-
mical redox processes compared to bulk materials
7–10
is
drawing attention.
Subnanometer pores/channels (SNPCs) are ubiquitous in
various rechargeable battery systems, yet the vital roles of
SNPCs have long been neglected. Based on the extensive reports
on electrode materials with SNPCs, such as metal oxides, metal
phosphides, metal nitrides/carbides, poly-anionic compounds,
Prussian blue, and carbon materials, the functions of SNPCs
can be summarized as (1) size/channel effects on fast ion
diffusion. With shortened ion-diffusion length in SNPCs,
fast diffusion of ions can be achieved, which is one of the
main indicators for the battery performance i.e., high-rate
capability.
11
In addition, SNPC-based materials can serve as
artificial protective shelves on the surface of the metal anodes,
which are capable of regulating uniform ion flux and prevent-
ing the side reactions of the metal anodes with the electrolyte
solvents, thus achieving high Coulombic efficiency and
dendrite-free stripping/platting processes.
12
(2) Abundant
active sites for enhanced ion storage kinetics and capacity.
Materials with a larger surface area usually have more active
sites, which can provide more storage sites for ions.
13–15
By
controlling the size, orderliness, and wettability of SNPCs, the
ion storage performance in batteries can be significantly
improved. For example, modifying the oxygen-containing func-
tional groups in SNPCs can change the surface and bulk
structures, conductivity, wettability, and reaction activity of
the material, thereby improving the electrochemical reaction
kinetics of sodium ion storage.
16,17
(3) Pathway regulation of
redox reactions. The SNPCs could enable selective transport of
various species (ions and/or molecules) through the size effect
of the pore/channel or electrostatic effect of the building
blocks, so as to avoid undesirable side reactions.
18
For
instance, the small molecule S species (S
2–4
) confined in the
SNPCs of the carbon matrix could achieve solid-state conver-
sion during the charge/discharge process, prohibiting the
formation of soluble polysulfides that lead to parasitic
reactions.
19
Other than the advantages of SNPCs in perfor-
mance improvement, some key properties of SNPCs and their
underlying relationships with reaction mechanisms are still
unclear and have been barely explored to date.
20–25
These include the synthesis of electrode materials with
uniform SNPCs and maximization of the effective utilization
of SNPCs for ion storage. Understanding the contribution of
SNPCs to ion storage will help us discover electrode materials
with higher capacity and reversibility. In addition, the internal
ion diffusion within the SNPCs and the ion diffusion from the
interface are the main factors affecting ion kinetics.
26,27
When
the resistance at the interface of two bulk materials signifi-
cantly increases, it can lead to a decrease in electrode potential
and low ion accessible area, especially at high currents, thereby
severely reducing performance. Furthermore, increasing the
thickness or mass loading of the electrodes will also severely
limit ion diffusion. Ideally, an efficient SNPC should achieve
fast electrolyte transport through ion diffusion and have good
electrolyte accessibility. Also, ions always exhibit close contact
within the SNPCs. In this way, the storage and transport of ions
in individual SNPCs are influenced by collective effects. How-
ever, these effects cannot be described by traditional channel
Yun-Xiao Wang
Yun-Xiao Wang currently holds
the position of Vice Dean at the
Institute of Energy Materials
Science (IEMS), University of
Shanghai for Science and
Technology. She received her
master’s degree from Xiamen
University in 2011 and obtained
her PhD degree from the
University of Wollongong in
2015. She was awarded a
Discovery Early Career
Researcher Award (DECRA) from
the Australian Research Council
in 2017. Her research features elaborate material engineering for
various emerging battery systems and other energy storage
technologies, with a particularly focus on developing room-
temperature sodium–sulfur batteries.
Guoxiu Wang
Guoxiu Wang is the Director of
the Centre for Clean Energy
Technology and a Distinguished
Professor at the University of
Technology Sydney (UTS),
Australia. Professor Wang is an
expert in materials chemistry,
electrochemistry, energy storage
and conversion, and battery
technologies. Currently, he
serves as an Associate Editor for
Electrochemical Energy Reviews
(Springer-Nature), and an
Associate Editor for Energy
Storage Materials (Elsevier). His research interests include
lithium-ion batteries, lithium-air batteries, sodium-ion batteries,
lithium–sulfur batteries, and electrocatalysis for hydrogen produc-
tion. Professor Wang has published more than 700 refereed journal
papers. His publications have attracted over 75 000 citations with
an h-index of 150 (Google Scholar). He has been recognised as a
highly cited researcher in both materials science and chemistry by
Web of Science/Clarivate Analytics.
Review Article Chem Soc Rev
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
This journal is © The Royal Society of Chemistry 2024 Chem. Soc. Rev.
theories. Importantly, only by studying the electrochemical
reactions within SNPCs at the atomic scale can these challenges
be fundamentally understood and addressed and their electro-
chemical performance be further improved.
To the best of our knowledge, no review has been reported
yet from the perspective of the electrochemical couplings of
SNPCs in rechargeable batteries. Herein, the SNPC families
have been systematically classified and summarized from the
architecture point of view, i.e., one-dimensional (1D) SNPCs,
two-dimensional (2D) SNPCs, and three-dimensional (3D)
SNPCs. These SNPCs embrace a wide range of materials includ-
ing carbonaceous materials, metal organic frameworks (MOFs),
covalent organic frameworks (COFs), porous organic cages
(POCs), MXenes, layered transition metal dichalcogenides
(TMDs), layered transition metal oxides (TMOs), transition
metal phosphates (TMPs), and various types of cathode materi-
als. Besides, diverse approaches to alter the structure and
building blocks of the SNPCs for anodes, cathodes, electrolytes,
and functional interlayers to enhance the electrochemical
performance of the rechargeable batteries have been intensively
reviewed. Furthermore, the gaps between practical application
and fundamental understanding of SNPCs in rechargeable
batteries have been carefully illustrated, while future research
directions and promising approaches are rationally suggested.
We hope that this review could shed light on the forefront of
this emerging yet underdeveloped field and stimulate further
exploration in this exciting area.
2. Scientific gaps and classifications of
SNPCs
2.1. Scientific gaps for understanding SNPCs in batteries
The last century has witnessed great achievements in manifold
characterization techniques for materials, enhancing the cap-
ability to identify elements, bonds, phase, and morphology of
the samples from micrometer to subnanometer scales. These
powerful tools are indispensable for the development of various
rechargeable batteries, and the progressive understanding of
the working mechanisms inside the batteries. The properties of
the materials can vary significantly with different sizes from
micro or nano to atomic scales, for example, the ratio of surface
atoms increases exponentially at subnanometer scales. The
angstrom-scale charge carriers exhibit intimate electrochemis-
try couplings with SNPCs in rechargeable batteries; however,
the capabilities of various characterization techniques to
understand these interactions in batteries at the subnanometer
scale are limited.
Electrochemical studies on SNPCs are expected to overcome
the limitations of previous nanopores and achieve higher
selectivity and electrochemical reaction efficiency.
28,29
Although
the application of SNPCs in batteries is still in the early stages,
it is undoubtedly a new topic in the development of next-
generation materials. At the SNPC-level, the operation of the
battery involves ion storage and ion transfer processes. They
typically form atomic and molecular-scale circuits. Exploring
the atomic-level assembly of cathode and anode electrode
materials, electrolytes (solid-state and quasi-solid-state), inter-
mediate layers/isolation layers, and other components from the
perspective of SNPCs to form composite electrode structures
with efficient ionic carriers and both ion and electron conduc-
tivity is of great significance for batteries. This is because the
size, shape, and spatial distribution of each component have a
decisive impact on the ion storage and diffusion, thereby
affecting the charge–discharge performance and rate capability
of the battery. As shown in Fig. 1, the research on SNPCs can
effectively upgrade various components of batteries, including
anodes, cathodes, solid-state electrolyte, as well as interlayers/
separators. Optimization of SNPCs can strengthen the interface
stability of anodes, especially for the SEI. An excellent SEI layer
with appropriate SNPCs can selectively transport ions, avoid
excessive SEI formation to reduce electrolyte consumption, and
reduce volume changes during the cycling process. In terms of
insertion-type cathodes, carefully designed SNPC materials can
effectively promote ion diffusion while maintaining material
structure stability, enhancing rate performance; for conversion-
type cathodes, structural stability and reaction kinetics are
crucial performance factors. By improvements in SNPCs, the
Fig. 1 Schematic illustration of the versatile advantages of SNPCs for various battery components.
Chem Soc Rev Review Article
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2024
active material can be better confined within SNPCs and
accelerate ion diffusion, which not only suppress the ‘‘shuttle
effect’’, but increase kinetic reactions. Moreover, well-designed
SNPCs for solid-state electrolytes can enhance their ionic con-
ductivity, and their strong selectivity advantage can effectively
shield unnecessary ions from passing through, avoiding the
generation of side reactions. Moreover, by leveraging the
advantages of high selectivity and excellent ionic conductivity
of SNPCs, ions can rapidly and orderly diffuse through the
separators/interlayers, avoiding uneven deposition and side
reactions.
However, designing and preparing efficient battery compo-
nent structures are challenging for the following reasons: (1)
there is still insufficient understanding of the dynamic
mechanisms of such complex battery systems; (2) the key ion
storage properties, as well as electronic and ionic transport
properties of many materials have not been determined or
there is uncertainty; and (3) involving multiphase and multi-
scale situations makes the problem more complex. Therefore,
studying batteries from the perspective of SNPCs helps to
comprehensively understand the electrochemical reaction
mechanisms in batteries and achieve high-performance bat-
teries with fast charge–discharge cycles, long life, excellent rate
stability, and high energy/power density. Nevertheless, a sys-
tematic summary of SNPCs and their electrochemical reactions
in batteries has not yet been collected. Therefore, this is the
first review to summarize and report on SNPCs and their
applications in batteries.
2.2. SNPC families
SNPC families represent a wide range of porous or layered
materials with a pore diameter or interlayer spacing of less than
1 nm, including various types of materials such as carbonac-
eous materials, inorganic transition metal compounds, porous
organic materials, and metal–organic materials. The carbonac-
eous materials include carbon nanotubes and layered graphene
analogues such as graphite, graphene oxide, reduced graphene
oxide, graphdiyne, and graphitic carbon nitrides (g-C
3
N
4
).
30
Although phosphorene, silicene, and germanene, are not car-
bonaceous materials, they exhibit layer structures and deliver
similar properties to that of graphene and can be reasonably
sorted as GAs.
31,32
Inorganic transition metal compounds
based SNPCs include transition metal oxides phosphates,
dichalcogenides, carbides, and nitrides.
33
Porous organic
material-based SNPCs include covalent organic frameworks,
porous organic cages, and metal–organic frameworks.
34
As
the shape, diameter, and building blocks of the SNPCs have a
crucial influence on the electrochemical performance of the
battery materials, logical classification and detailed discussion
are essentially required. Therefore, we tentatively classify the
SNPCs from the architecture point of view into three categories,
i.e., liner shape one-dimensional (1D) SNPCs, channel shape
two-dimensional (2D) SNPCs, and interconnected three-
dimensional (3D) SNPCs. The typical representatives, distinc-
tive properties, and potential applications of the SNPC families
will be comprehensively illustrated in the following sections.
2.2.1. 1D SNPCs
Carbon nanotubes (CNTs). As presented in Fig. 2, CNTs are
cylinders of one or more layers of graphene with open or closed
ends, which are usually divided into three types according to
the number of the rolling graphitic sheets, i.e., single-walled
CNTs (SWNTs), double-walled CNTs (DWCNTs), and multi-
walled carbon nanotubes (MWNTs).
35,36
Besides, other types
of CNTs with non-perfection structures also deserve wide
attention, since the doped heteroatoms or defective sites in
CNT structures could significantly alter the properties of the
CNTs.
37,38
CNTs can exhibit either metallic or semiconductor
properties, depending on the direction of the graphitic sheets,
Fig. 2 Illustration of representative 1D SNPC materials. (a) Representative materials of CNTs. (b) Representative materials of COFs. (c) Representative
materials of MOFs. (d) Representative materials of TMOs/TMPs.
Review Article Chem Soc Rev
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
This journal is © The Royal Society of Chemistry 2024 Chem. Soc. Rev.
e.g. zigzag,
39
armchair,
40
or chiral configurations.
41
SWCNTs
exhibit excellent chemical stability and provide efficient
chemical shielding, which could effectively protect the encap-
sulated species from external influences.
42
For instance, S has
been confined in the 1D SNPCs of the SWCNTs with a pore
diameter of 0.99 nm, which could facilitate favourable solid-
state reactions in glyme-based electrolytes with inhibited S
dissolution.
43
Other distinctive properties including good
mechanical stability, excellent electrical conductivity, and tune-
able SNPC size, render them promising as backbones for
various composite materials in battery applications.
Covalent organic frameworks (COFs). COFs are highly crystal-
line, covalently linked, periodically extended organic networks
consisting of only light elements (i.e., H, B, C, N, and O).
44
They
exhibit high structural, chemical, and thermal stability, as well
as a rich porous structure for a wide variety of applications such
as gas storage, separation, catalysis, and sensors.
45–48
Notably,
there is a growing interest in utilizing COFs in the realm of
electrochemical energy storage, including those with 1D SNPCs
as presented in Fig. 2, such as COF-1, COF-5, COF-6, ACOF-1,
and JUC-505.
49
The pore size of COFs ranges from subnan-
ometer to nanometer scales, which can be rationally engineered
by altering the building blocks.
50–52
The utilization of COFs
with 1D SNPCs could facilitate fast ionic transport through
SNPCs, while blocking the undesirable species with a diameter
larger than the SNPCs. For instance, a redox-active, crystalline
COF has been reported as a cathode material for LIBs with
excellent rate capability and good cyclability, as a result of
enhanced ion transportation derived from the SNPCs of the
COF structures.
53
COFs possess distinctive qualities, including
their inherent insolubility in electrolytes, abundant porosity,
ordered open SNPCs that favor ion transportation, and p-
conjugated frameworks that enhance charge transfer.
54–56
Con-
sequently, the prospects of COF-based electrode materials for
rechargeable batteries are undeniably promising.
Metal–organic frameworks (MOFs). MOFs are porous
materials consisting of metal ions or clusters coordinated
with organic ligands, which build up extended crystalline
structures with excellent structural stability upon
intercalation/extraction of guest molecules.
57–59
There are
plenty of MOFs with 1D SNPCs in diverse MOF families,
including ZIF-8, ZIF-69, CPO-27, MIL-53, MOF-74, etc.,as
presented in Fig. 2. In the field of rechargeable batteries, the
SNPCs in MOFs can offer a large surface area and adaptable
pore size, which facilitates quick diffusion of cations.
60–63
These merits make them ideal for intercalation-type
electrodes. Additionally, MOFs with SNPCs possess
confinement effects to effectively trap S species, enabling
their utilization as host materials for conversion-type
electrodes in metal–sulfur batteries. Furthermore, the well-
ordered channels and unique features of MOFs with SNPCs
can enhance the uniformity of cation plating during the
charging and discharging process, which could result in the
formation of a stable interface layer and enhance cycling
stability.
64,65
Consequently, MOFs with SNPCs are also
attractive for applications in solid-state electrolytes (SSEs) and
artificial interfaces. Overall, MOFs with SNPCs have multiple
features that give them enormous potential for various kinds of
batteries.
Transition metal oxides (TMOs) and transition metal phos-
phates (TMPs). Ionic channels are indispensable for various
cathode materials, and the size of these channels falls within
the realm of SNPCs. TMOs and TMPs are among the most
representative cathode materials extensively explored for alkali
metal ion batteries, of which olivine and tunnel structured
cathodes exhibit rich 1D SNPCs.
66,67
As presented in Fig. 2, the
cations of the cathodes are inserted in the 1D SNPCs build by
the M–O (M = transition metals) and/or P–O polyhedrons,
which is the case for LiFePO
4
, NaFePO
4
,Na
0.44
MnO
2
,Li
2
FeSiO
4
,
and KVOPO
4
. The shape and size of the SNPCs in the cathodes
could be altered using various approaches to modify ion
transportation inside the SNPCs, which will be discussed in
detail in the following sections.
2.1.2. 2D SNPCs. Since the discovery of graphene by Geim
and Novoselov in 2004, 2D materials have surged in various
fields of applications, especially in rechargeable batteries.
68–71
Diverse 2D materials with 2D SNPCs have been successfully
fabricated as promising candidates for rechargeable batteries,
including metal sulfides,
72
metal oxides,
73
metal nitrides,
74
metal carbides,
75
metallenes,
76
etc. (Fig. 3) In comparison to
other materials, 2D materials exhibit many unique properties,
such as excellent conductivity, mechanical flexibility, uniform
electronic state and lattice planes, desirable SNPC structures,
etc., making them potential electrode materials for various
rechargeable batteries.
2,77–79
Graphene analogues (GAs). Graphene is the most prestigious
2D material with a single layer of sp
2
carbon atoms arranged in
a hexagonal lattice structure.
80
Graphene possesses many
unique properties, including a high surface area, high modulus
of elasticity, good mechanical stability, and excellent electrical
conductivity.
81
It is worth noting that single-layer graphene is
relatively scarce and difficult to obtain; most of the research has
been carried out on few-layer graphene, which is composed of
several parallel graphene layers with weak interlayer van der
Waals force.
82
The size of the channel between the graphene
layers falls into the region of SNPCs, which could allow for the
intercalation of various metal ions as well as solvents. For
example, graphene oxide laminates were fabricated by Abra-
ham et al., which exhibit distinctive SNPCs with well-defined
interlayer spacing ranging from 6.4 to 9.8 Å for ion
intercalation.
83
On the other hand, generating defects in the
graphene layers (holey graphene) could facilitate ion transport
across the graphene layers. For instance, holey graphene with
SNPCs of 0.5 to 0.8 nm have been produced through high-
temperature etching,
84
and ion permeable SNPCs can be engi-
neered on the graphene layer through plasma etching.
85
Soon
after the discovery of graphene, the exploration of 2D materials
has been extended to a wide variety of other graphene
Chem Soc Rev Review Article
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2024
analogues (GAs) with similar 2D layered structures but different
chemical compositions, including graphene oxide (GO),
reduced graphene oxide (RGO), graphdiyne, silicene, germa-
nene, g-C
3
N
4
,etc. These GAs with distinctive 2D layered struc-
tures and rich SNPCs have been widely used in various
applications including catalysis, superconductors, sensors, as
well as energy storage and conversion devices including various
batteries.
MXenes. MXenes stand for a wide variety of 2D metal
carbides, nitrides, or carbonitrides with an odd number of
layers in which the metal layers (M) are sandwiched between
carbon or nitrogen (X) layers.
86
Since its first discovery by Yury
et al. in 2011, their unique features, including electrical con-
ductivity, mechanical stability, and thermal stability, render
them promising for various applications in supercapacitors,
batteries, supercapacitors, catalysis, membranes, and
sensors.
87
Mxenes exhibit the general formula of M
n+1
X
n
T
x
,
where M is the metal, X represents C or N, and T refers to the
terminal groups such as –H, –F, and –OH, which are obtained
by selectively etching the A layer (usually the Al layer) from
pristine MAX precursors. After ten years of development,
MXenes have been extended from Ti
3
C
2
to a large family of
metal carbides/nitrides with metals such as Ti, Sc, Zr, Hf, V, Nv,
Ta, Mo, Cr, Nb, or Ta.
6,86
The interlayer space of MXenes ranges
from about 3.5–5 Å, making them suitable SNPCs for Li and Na
ions storage.
6,88,89
Besides, enriched functional groups in the
SNPCs of MXenes could efficiently trap polysulfides, thereby
inhibiting the shuttle effect in metal–S batteries.
90–92
MXenes
have shown great potential in various applications when
coupled with a diverse range of other materials.
Transition metal dichalcogenides (TMDs). TMDs are layered
materials with a general chemical formula of MX
2
, where M
stands for transitional metals and X represents chalcogen
elements of S, Se, or Te.
93–95
The properties of TMDs can be
varied from insulators (HfS
2
) to semiconductors (MoS
2
and
WS
2
), semimetals (WTe
2
and TiSe
2
), and metallic (NbS
2
and
VSe
2
), rendering them capable for a variety of applications.
96
2D
TMDs always feature multi-crystalline structures.
97
Taking
MoS
2
as an example, it has three crystal structures, namely,
2H (hexagonal), 1T (trigonal), and 3R (rhombohedral), each
characterized by distinctive coordination models and staking
orders. Different from graphene, 2D TMDs can expose more
active sites, further improving the electrochemical performance
of TMDs.
74,98,99
In addition, the interlayer spacing of 2D TMDs
at the subnanometer level has strong covalent bonds within the
layer and weak van der Waals forces between layers, providing
an ideal space for intercalation of ions with small ionic radius,
such as Li or Na ions.
2,100–102
Layered TMOs. Due to high structural stability, low material
cost, large specific capacity, high ionic conductivity, good
environmental benignity, and feasible synthesis conditions,
layered TMOs have been developed as one of the most popular
candidates of cathodes for LIBs and SIBs.
103–106
Generally, the
structure of layered TMOs consists of alternating MeO
2
octahe-
dra layers and alkali metal layers, which can be sorted as O- or
P-type depending on the coordination environment of the alkali
metal ions.
107,108
The interlayer spacings of layered TMOs are
in the range of SNPCs, facilitating quick 2D ion transportation
through the SNPCs and resulting in excellent rate capabilities.
2.1.3. 3D SNPCs
MOFs. 1D SNPCs are ubiquitous in various MOFs. For MOFs
with 1D SNPCs at intersecting directions, 1D SNPCs would
interconnect with each other to generate 3D SNPCs.
109
For
instance, as illustrated in Fig. 4, Prussian blue analogues
(PBAs), one of the earliest discovered MOF materials, exhibit
1D SNPCs at three intersecting directions, generating a 3D
Fig. 3 Illustration of representative 2D SNPC materials. (a) Representative materials of GAs. (b) Representative materials of MXenes. (c) Representative
materials of TMDs. (d) Representative materials of TMOs.
Review Article Chem Soc Rev
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
This journal is © The Royal Society of Chemistry 2024 Chem. Soc. Rev.
SNPC network.
110
These PBAs with 3D SNPCs can facilitate
quick diffusion of metal ions, together with abundant redox-
active sites and good structural stability, making them an ideal
choice for alkali metal ion batteries. Similarly, other MOFs with
3D SNPCs, such as MIL-125, UIO-66, MOF-177, and HKUST-1,
have also been widely used in various battery systems.
111–113
The advantages of utilizing these MOFs with 3D SNPCs for
batteries will be discussed in detail in the following sections.
COFs. Similar to MOFs with 3D SNPCs, the interconnection
of 1D SNPCs in the COFs results in 3D SNPCs. The COFs with
3D SNPCs exhibit not only cross-linked 3D electrically conduc-
tive frameworks but also interconnected 3D ion permeable
pathways. Therefore, they are promising candidates for active
materials or coating layers in various battery systems.
114–116
Those frequently reported COFs with 3D SNPCs including
COF-102, COF-105, COF-108, COF-300, CTF-1, etc.
117
Porous organic cages (POCs). POCs are a series of micro-
porous organic molecules with intrinsic, guest accessible
cages.
118
They are built through covalent bonds involving
non-metal elements, such as carbon–carbon or carbon-
heteroatoms, imines, boronic esters, amides, etc. Compared
to other porous frameworks, POCs exhibit many merits includ-
ing large surface areas, pore volumes as well as open and
flexible pores, including tuneable SNPCs.
28,119,120
In addition,
their discrete molecular structure provides excellent solubility
in common solvents with high solution dispersion and proces-
sability, which is a rare feature that cannot be achieved in
insoluble extended porous frameworks such as MOFs and
COFs.
121,122
A wide variety of POCs including CPOC-102,
CPOC-104, CPOC-302, RCC1, and CC-8 have been used in
various battery systems, whose functions and advantages would
be detailed in the following sections.
122,123
TMOs and TMPs. Besides, 1D and 2D SNPCs in olivine and
layered type cathodes, 3D SNPCs have also been extensively
reported in various cathodes such as Na
3
V
2
(PO
4
)
3
, which has
been widely known as a NASICON (sodium super ionic con-
ductor) type cathode. These NASICON cathodes generally
undergo topotactic insertion/extraction Na
+
with small volume
changes and minimal structural rearrangements, featuring
outstanding long-term cycling stability.
124,125
Plenty of cathode
materials with 3D SNPCs have been widely investigated to
exhibit high ionic conductivity in the cathode matrix, which
favours high rate performance of the cathodes. Various other
cathode materials with 3D SNPCs, such as LiMn
2
O
4
and
Na
2
FeP
2
O
7
,
126
will also be discussed in the following sections.
It should be noted that simply utilizing one kind of porous
material is not the case in most scenarios; instead, hetero-
structures or composite materials that merge the merits
of various components have been frequently reported.
127–129
In particular, the built-in electric field at the heterostructure
interface can promote rapid transport of ions and
electrons.
130,131
Heterostructures composed of adsorption com-
ponents and catalytic components can combine two functions
to achieve synergetic effects.
132,133
For instance, higher energy
density can be achieved in cathode materials by carbon-coating
with three-dimensional sodium diffusion channels.
134
More
rational designs of heterostructures and hybrid materials will
be illustrated in detail in the following discussions.
2.3. Unique physicochemical properties of SNPCs
As the size of the pore/channel decreases to the subnanometer
level, which is within the scale range for molecules, ions, and
atoms, unique physicochemical properties differing from those
of the bulk materials would emerge.
29,135,136
Some extraordin-
ary properties that are impossible in the view of classic physics
would take place, which makes SNPCs a totally new world of
Fig. 4 Illustration of representative 2D SNPC materials. (a) Representative structures of MOFs. (b) Representative structures of COFs. (c) Representative
structures of POCs. (d) Representative structures of TMOs/TMPs.
Chem Soc Rev Review Article
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2024
exciting wonders. Generally, the ionic conductivity could be
defined as:
k¼GL
S(1)
where Gis the conductance of an ionic channel, Sis its cross-
sectional area and Lis the length. But in fact, the ionic
conductivity of nanochannels is quite complex. The diffusion
of monovalent ions confined within a pore with a width like the
ion diameter can be expressed as:
J¼DrrDrGrCDr
rmax rS
rrS (2)
where Dis the diffusion coefficient of the ion (for simplicity, we
will use the same D=D
), r
represents the ion density, c=r
+
r
represents the charge (in units of the elementary charge),
and r
S
represents the total ion density. Additionally, r
max
is the
total ion density at close packing. Gis a parameter that
characterizes the screening of ion–ion electrostatic interac-
tions. To realize G, first, we need to consider the exponential
screening of ion–ion electrostatic interactions caused by the
pores. In addition, we adopt the Borukhov–Andelman–Orland
entropy, S[r
].
137
The voltage-dependent ‘‘external field’’
includes the electrochemical potential of ions, solvation energy,
and van der Waals and image forces between ions and pore
walls. The voltage-dependent ‘‘external field’’ is composed of
the electrochemical potential of ions, solvation energy, and van
der Waals and image forces between ions and pore walls. The
external field controls the equilibrium ion density inside the
pore, and besides participating in dynamics through initial and
boundary conditions, it does not participate in other processes.
Dynamics is defined by the equations of continuity: q
t
r
=
q
w
J
.J
=Gr
q
w
(dF/r
), where GG
=D
/k
B
Tis a
phenomenological mobility parameter, Dis the diffusion con-
stant, and we assume it is independent of pore width, voltage,
and density; k
B
is the Boltzmann constant. Thus, plugging the
free energy Finto the continuity equation yields eqn (2),
138
G¼4LBRCSðsin pn=2Þ2
nK1pnRC=LðÞ (3)
From eqn (3), it can be determined that Gis constrained by
many factors, including ion density, charge within the channel,
electrostatic forces, gradient of ion density along the pore,
collective properties, etc. By changing the above parameters,
selective transportation of specific ions can be achieved. Addi-
tionally, confining restricted species at the molecular scale with
distinctly different properties is another appealing feature. For
battery couplings, SNPCs also possess unique properties, which
can be reasonably classified into six categories size effect,
electrostatic effect, confinements effect, suppression effect,
buffering effect, and quantum effect (Fig. 5).
Size effect. The size effect plays an important role in mass
transfer. Due to the small size of SNPCs, smaller ions can be
allowed to diffuse from one side to the other side while ions
with larger size will be inhibited from entering the SNPCs. This
function has been widely used in rechargeable batteries, for
example, artificial SEIs with abundant ion diffusion channels
have been designed, which often only allow smaller ions such
as Li
+
,Na
+
,K
+
,etc. to pass through while inhibiting the passage
of other larger ions.
139–142
This is beneficial for supporting high
ion flux for homogeneous metal stripping/plating, and also
helps to avoid adverse reactions between larger solvent mole-
cules and ions which are in directly contact with the electrode.
Therefore, by adjusting and improving the specific SNPCs of
the materials, potential barriers to mass transfer and storage
can be reduced.
Electrostatic effect. Mass transport of various species
through the SNPCs via diffusion or advection caused by ther-
mal motion or electric force is essential for electrical chemical
reactions in the batteries; one of the most important factors
impacting the mass transport is the electrostatic effect.
143–145
The electrostatic effect of SNPCs can be easily conceived, as the
building blocks would attract species with opposite charges
while repelling those with the same charges.
146–148
It should be
noted that the electrostatic effect of SNPCs is generally induced
by partial charges caused by the asymmetric distribution of
electrons in the chemical bonds, rather than bare charge which
tends to be unstable. For instance, Lewis acidic sites in the
SNPCs of COFs and MOFs can complex with anions in the
electrolyte, which is beneficial for the quick transport of
cations.
149
The surface of graphene oxide (GO) membranes
with SNPCs can be engineered with positive or negative
charges, which can repel ions with same charges while allowing
the transmission of ions with opposite charges.
150
Therefore,
the building blocks of the SNPCs can be rationally designed to
achieve selective ion transportation through electrostatic
effects to screen ions with opposite charges.
Confinements effect. Physical confinement is essential for
electrode materials that have poor electric conductivity or
exhibit undesirable side reactions with the electrolytes, which
is essential for sulfur (S) cathodes since they exhibit both these
problems.
151,152
Whereas the confinement effect of SNPCs for S
cathodes is significantly different from that in the larger pores,
where the S molecules are reduced to short chain small
molecules (S
2–4
) instead of long chain S
4–8
species. For example,
Wan et al. demonstrated that S
2–4
molecules can be accommo-
dated in conductive carbon matrix with SNPCs (B0.5 nm),
instead of cyclo-S
8
which exhibits a size larger than 0.5 nm in
at least two dimensions.
153
As such, the formation and dissolu-
tion of long chain polysulfides can be avoided, so as to achieve
good cycling stability and superior rate capability. Furthermore,
it has been evidenced that S encapsulated in the narrow-
diameter SWCNTs with SNPCs undergoes solid-state reactions,
since the Li ions are de-solvated before they enter the SWCNTs
as a result of the confinement effect of SNPCs.
43
Similarly, the
confinement effect of SNPCs can be extended to other electrode
materials to prevent them from side reactions with electrolytes.
Suppression effect. SNPCs often exhibit suppression effects
as a result of their good mass transport ability and mechanical
stability, which could allow for ions uniformly transmit
through the SNPCs while suppressing short circuits caused by
dendrite propagation. This property has been widely applied in
Review Article Chem Soc Rev
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
This journal is © The Royal Society of Chemistry 2024 Chem. Soc. Rev.
various battery systems. For example, common SPEs (such as
PEO) are difficult to effectively suppress lithium dendrites due
to their low mechanical strength, which can easily lead to
battery failure due to short circuits. However, adding ceramic
fillers (such as LLZAO) with rich SNPC structures and good
mechanical properties to polymer electrolytes can not only
increase the ion transport pathways and boost the speed of
ion migration, but also effectively increase the mechanical
strength of the electrolyte which is beneficial for suppressing
lithium dendrites.
154,155
Buffer effect. When the dimensions of the material (such as
the aperture or thickness) are close to the size of a single crystal
cell, nanoscale materials exhibit some polymer-like properties
due to their similarity in size to polymer chains. When the
material dimensions are larger than the nanoscale, they are
typically rigid and linear, but when the dimensions are close to
or smaller than the nanoscale, they become flexible or curved,
like polymer chains.
156,157
In the diffusion process of ions, the
conformation of SNPCs changes under shear stress, resulting in
shear stress being disproportionate to shear rate and exhibiting
a good buffering effect. For instance, a block polymer was
designed as an interlayer to buffer tbe cathode volume change
during discharging and charging processes.
158
Therefore, this
characteristic can be applied to non-viscous electrodes and
solid-state electrolytes in batteries to achieve good mechanical
performance, prevent material fracture and resist volume
expansion effects.
158,159
Quantum effect. The quantum effect refers to a series of
distinctive phenomena observed when the size of transported
particle is close to the Ångstrom scale range, which follow the
rules of quantum mechanics and cannot be explained by
classical physics. From the perspective of classical thermody-
namics, the chemical selectivity mass transfer through nano-
scale channels should be very slow: this limitation is predicted
by the Hagen–Poiseuille equation, as the traditional laminar
flow has zero flow velocity at the pore wall.
160
One of the
quantum effects is tunnelling, which refers to the penetration
of a particle getting through an energy barrier that is higher
than the particle’s kinetic energy.
161,162
The principle of tunnel-
ling has led to the development of the scanning tunnelling
microscope (STM), which has significantly advanced scientific
research to atomic scale resolution.
163
Another quantum effect
for SNPCs is quantum confinement, where the movement of
particles will be restricted within the SNPCs, leading to the
existence of particles only in the form of quantum states when
the size of particles approaches the size of the SNPCs. For
Fig. 5 Schematic illustration of the unique physicochemical properties of SNPCs.
Chem Soc Rev Review Article
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2024
instance, quantum confinement occurs between the interlayer
space of van der Waals channels.
164
In terms of ions diffusion
in SNPCs, however, the rapid ion transport is precisely caused
by quantum flow, resulting in superfast ion fluid states. A
quantum ions fluid can orderly arrange ions, greatly reducing
the ion diffusion energy barrier and achieving highly efficient
and selective ion diffusion.
3. Electrochemical couplings in SNPCs
As mentioned above, electrochemical coupling of SNPCs in
batteries is ubiquitous. To understand the relationship
between SNPCs and the electrochemical performance, we have
summarized the size of SNPCs and various other electrochemi-
cal parameters in batteries (Tables 1–4). We have tentatively
sorted the specific capacities (for cathodes and anodes) and the
ionic conductivities (for electrolytes and separators/interlayers)
according to the size of the SNPCs, to understand the influence
of SNPCs on the electrochemical performance of the batteries.
As shown in Fig. 6a, capacities of transition metal oxides,
polyanions, PBAs, and polymers exhibit not directly relation-
ship with SNPC size. The possible reason is that these materials
have enough pores to provide ion diffusion and storage, and
the reaction kinetics are all not slow. Instead, their perfor-
mance is more likely to be influenced by electro-conductivity
rather than ionic conductivity. Nevertheless, it is obvious that
with the increase in pore size of S/Se cathodes and COFs
(Fig. 6b), the performance significantly decreases. This
indicates that the reaction kinetics are one of the key factors
determining their capacities. As for the anodic materials in
Fig. 6c, the pore size of phosphides, sulfides, and polymers,
does not directly affect their ion storage. However, for MXenes,
graphite-based materials, and hard carbons shown in Fig. 6d,
the size of the pores is closely related to ion storage. For
MXenes, increasing the size of the pores will lead to a decrease
in ion storage performance, while for graphite-based materials
and hard carbons, the capacity will increase with the increase
in pore size. This indicates that the regulation of the pore size
will affect the ion storage capacity of these carbon-based anodic
materials. For solid-state electrolytes and quasi-solid-state elec-
trolytes, it is commonly known that ionic conductivity can
impact their performance. According to Fig. 3e, however, the
ionic conductivity of some materials does not show a linear
relationship with SNPC size. It may be because when polymers
and MOFs are used as electrolytes, they are usually modified. As
described in eqn (3), changing their electrostatic force, ion
concentration difference, and other conditions can greatly alter
the ion conductivity. Therefore, these materials do not exhibit a
linear relationship with size of SNPCs. However, for some
materials that are difficult to modify, such as NASICONs and
sulfides, their ionic conductivity exhibit a linear relationship
with SNPC size. Thus, as the pore size increases, the ion
conductivity also increases in Fig. 6f. For interlayer/separators,
their research mainly focuses on polymers, oxides, and MOFs.
As illustrated in Fig. 6g, there is no clear correlation between
the ionic conductivity and SNPC size. Nonetheless, it is obvious
that the ionic conductivity of MOFs is affected by the size
of SNPC.
The analysis of Fig. 6 shows that the determining factors of
battery performance are not only the reaction kinetics but also
the redox reaction activity and electronic conductivity of the
materials. Ion diffusion is influenced by various factors, which
can alter its kinetic activity and result in differences in battery
performance. To help readers better understand the main role
of SNPCs in rechargeable batteries, we have compiled and
summarized some representative studies.
3.1. Electrochemical couplings of cathode materials with
SNPCs
During the electrochemical reactions of rechargeable batteries,
the cathode materials coupling with SNPCs serve as the hosts
for cations. The redox reactions during charging and dischar-
ging processes always cause expansion and contraction of
cathode materials, leading to undesired structure distortion,
capacity fading, or even battery failure. The design of SNPCs in
cathode materials can effectively facilitate ion diffusion and
buffer volume changes, ensuring stable performance in
rechargeable batteries. In this section, we outline the develop-
ment and achievement of SNPC design in cathode materials,
including intercalation-type and conversion-type materials
(Fig. 7).
3.1.1. Intercalation-type cathode materials
3.1.1.1. Tuning SNPCs in layered oxide cathodes. Layered
oxides with SNPCs, which have the chemical formula AMO
2
(A normally represents Li, Na, K, etc.; M represents transition
metals), are the most popular and well-developed cathode
materials under massive production in the current battery
industry.
543
The first commercial intercalation-type layered
oxide cathode material (LiCoO
2
) with SNPCs was introduced
by Goodenough et al. This cathode material attracted global
attention due to its high specific capacity (B140 mA h g
1
) and
super high working voltage in organic electrolytes, as well as its
excellent SNPC structure for excellent reversible Li
+
insertion
and extraction.
544,545
Due to the high cost and geographic
distribution issues of cobalt resources, many alternative layered
oxides with abundant and low-cost transition metals, such as
Ni, Mn, and Fe, were developed to form the layered LiMO
2
structure.
546–548
Moreover, Li was also replaced with Na or K to
form layered NaMO
2
and KMO
2
, which further reduces the
cathode cost.
549,550
However, with the decrease in the cost, the
challenges that affect the structure stabilities, energy density
and cycling performances of layered oxide cathode materials
have become more prominent.
551
The electrochemical energy storage mechanisms of the
layered oxides exploit the SNPCs (B0.4 to 0.7 nm) between
the adjacent MO
2
layers with edge-shared octahedral structures
to allow the ‘A’ ion insertion/extraction during the cycling
(Fig. 8a).
552–555
This mechanism inevitably subjects the MO
2
layers to structural distortion during repeated cation insertion/
extraction for long-term cycling, leading to structure degrada-
tion and capacity fading. To maintain the performance of
Review Article Chem Soc Rev
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
This journal is © The Royal Society of Chemistry 2024 Chem. Soc. Rev.
Table 1 Summary of the cathode materials with SNPCs (theoretical = T)
Materials Battery types
Theoretical
capacity
(mA h g
1
)SNPC
size (nm) Average
potential (V)
Capacity (mA h g
1
)@
current density (mA g
1
)@
cycle number Ref.
LiCoO
2
LIBs 274 4 123@50@10 165
Oxygen vacancy-Li[Ni
0.2
Li
0.2
Mn
0.6
]O
2
LIBs 315 3.8 121@60@10
Oxygen vacancy-Li
2
MnO
3
LIBs 230 2.8 59.5@10@50 166
LiCoO
2
LIBs 274 4 116@100@20 167, 168
LiNi
0.8
Mn
0.1
Co
0.1
O
2
LIBs 190 210@100@50 169
LiNi
0.8
Co
0.15
Al
0.05
O
2
LIBs 180 0.47 3.8 170@180@100 170
LiNiO
2
LIBs 275 3.8–3.9 140@275@100 171
LiNi
0.9
Co
0.1x
Ti
0.03
O
2
LIBs 275 3.8–3.9 180@275@50
LiNi
0.9
Co
0.08
Ti
0.02
O
2
LIBs 275 3.8–3.9 120@275@50
LiNi
0.9
Co
0.06
Ti
0.04
O
2
LIBs 275 3.8–3.9 190@275@50
Li
2
MnO
3
LIBs 458 0.47 4 140@45@20@10 172, 173
LiNi
0.8
Co
0.15
Al
0.05
O
2
LIBs 279 0.47 3.8 155@280@100 174
LiNi
0.6
Co
0.2
Mn
0.2
O
2
LIBs 279 3.6 170@280@100 175
Li
1.2
Mn
0.54
Ni
0.13
Co
0.13
O
2
LIBs 378 3.6 177@280@200 176
Li
1.13
Ni
0.30
Mn
0.57
O
2
LIBs 343 0.79 3.0 85@340@500 177
Li
1.2
Ni
0.17
Mn
0.56
Co
0.07
O LIBs 465 3.8 172@1400@200 178
LiNi
1y
Co
y
O
2
LIBs 276 3.7 148.4@27@20 179
LiNi
1x
Co
x
O
2
LIBs 276 3.8 175@137@50 180
Li
12x
Ni
0.85
Co
0.15
O
2
LIBs 180 181
LiNi
1/3
Co
1/3
Mn
1/3
O
2
LIBs 279 3.8 186@28@50 175
LiNi
0.7
Mn
x
Co
0.3–x
O
2
LIBs 160 182
LiNi
x
Co
y
Mn
z
O
2
LIBs 200 0.47 3.8 187@100@100 183
Z-doped LiNi
0.6
Co
0.2
Mn
0.2
O
2
LIBs 160 3.8 130@53.3@50 184
F-LNCM LIBs 200 0.55 3.75 153.4@200@200 185
5%P@LLO LIBs 250 3.5 200@250@100 186
B@LNMO LIBs 200 0.48 3.6 200@100@300 187
LiNi
0.6
Co
0.2
Mn
0.2
O
2
LIBs 180 0.47 3.8 141@180@200 188
LiNi
0.8
Co
0.15
Al
0.05
O
2
LIBs 180 3.8 140@36@1500 189
LiNi
0.7
Co
0.15
Mn
0.15
O
2
LIBs 180 3.8 130@36@1500
Mo-doped LiNi
0.5
Mn
0.5
O
2
LIBs 180 0.47 3.8 105@50@100 190
Nb modified NCM811 LIBs 200 0.47 3.8 182@66@250 191
LiNi
0.6
Co
0.2
Mn
0.2
O
2
LIBs 180 0.47 3.9 183.3@180@80 192
Ni-rich LiNi
0.8
Co
0.1
Mn
0.1
O
2
LIBs 180 0.47 3.8 178@180@100 193
Ti-doped Ni-rich LiNi
0.8
Co
0.1
Mn
0.1
O
2
LIBs 180 0.47 3.8 130@180@150 194
Ni-rich LiNi
1xy
Mn
x
Co
y
O
2
LIBs 190 3.8 140@62.7@600 195
LiCoMnO
4
LIBs 145 4.8 85@14@50 196
LiFe
0.5
Mn
1.5
O
4
LIBs 147 4.2 80@14@50 197
LiFe
0.1
Mn
1.9
O
4
LIBs 147 89@14@50
LiFe
0.3
Mn
1.7
O
4
LIBs 147 70@14@50
LiCu
0.5
Mn
1.5
O
4
LIBs 145 2.6 47@14@14 198
LiNi
0.25
Cu
0.25
Mn
1.5
O
4
LIBs 145 2.2 57@14@8
LiNi
0.5
Mn
1.5
O
4
LIBs 145 2.7 48@14@20
NaFeO
2
SIBs 241.8 0.53 2.8 48@200@100 199
NaNi
0.5
Mn
0.5
O
2
/Super P SIBs 240 0.51 2.8 142@12@100 200
NaNi
0.5
Mn
0.5
O
2
/CNT SIBs 240 0.51 2.8 127@12@100
NaNi
0.5
Mn
0.5
O
2
SIBs 240 0.54 2.9 74@240@200 201
Na
2/3
Ni
1/3
Mn
2/3
O
2
SIBs 173 3.2 52@346@500 106
Na
3/4
(Li
1/4
Mn
3/4
)O
2
SIBs 291 2.6 112@150@100 108
NaNi
0.5
Mn
0.2
Ti
0.3
O
2
SIBs 240 0.54 3.0 93@240@200 201
NaLi
0.1
Ni
0.35
Mn
0.55
O
2
SIBs 268 3.1 110@12@100 202
Na
1.2
Mn
0.4
Ir
0.4
O
2
SIBs 203 2.5 68@20@50 203
NaFe
0.55
Mn
0.45
O
2
SIBs 0.55 3.2 32@240@100 204
Na
0.75
Ni
0.82
Co
0.12
Mn
0.06
O
2
SIBs 200 0.55 2.8 80@200@400 205
Na
0.9
Ca
0.05
Ni
1/3
Fe
1/3
Mn
1/3
O
2
SIBs 130 3.0 102@130@200 206
Na[Li
0.05
(Ni
0.25
Fe
0.25
Mn
0.5
)
0.95
]O
2
SIBs 170 3.4 120@85@40 207
Na
0.85
Li
0.10
Ni
0.175
Mn
0.525
Fe
0.2
O
2
SIBs 150 3.5 115@150@140 208
Na[NiCoMnTi]
1/4
O
2
SIBs 120 0.54 3.0 78@600@400 209
NaMn
0.48
Ni
0.2
Fe
0.3
Mg
0.02
O
2
SIBs 240 3.0 100@24@100 210
NaCr
1/3
Fe
1/3
Mn
1/3
O
2
SIBs 200 2.7 102@5@12 211
Na[Cu
0.22
Fe
0.30
Mn
0.48
]O
2
SIBs 240 3.3 100@10@100 212
TiS
2
LIBs 239 0.57 2.25 200@240@100 213
K
0.3
MnO
2
PIBs 82 0.62 2.8 74@28@50 214
K
0.6
CoO
2
PIBs 80 2.5 44@100@120 215
K
0.7
Fe
0.5
Mn
0.5
O
2
PIBs 164 0.69 2.0 70@500@200 216
Na
0.7
Fe
0.5
Mn
0.5
O
2
SIBs 182 0.56 68@500@100 211
Li
0.7
Fe
0.5
Mn
0.5
O
2
LIBs 204 0.47 20@500@100
Na
0.7
CoO
2
SIBs 175 0.54 2.6 60@30@40 217, 218
100@175@100
Chem Soc Rev Review Article
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2024
Table 1 (continued )
Materials Battery types
Theoretical
capacity
(mA h g
1
)SNPC
size (nm) Average
potential (V)
Capacity (mA h g
1
)@
current density (mA g
1
)@
cycle number Ref.
NaNiO
2
SIBs 237 0.53 2.9 100@23.7@20 219
NaFeO
2
SIBs 244 0.56 3.2 68@12@30 220
NaCrO
2
SIBs 253 0.53 3.0 121@126@300 221
NaVO
2
SIBs 255 0.58 1.6 120@52@15 222
Na
0.7
VO
2
SIBs 190 0.58 1.6 100@38@10
NaNi
0.5
Mn
0.5
O
2
SIBs 240 0.56 3.5 138@24@12 223
Na
0.9
Cr
0.9
Ru
0.1
O
2
SIBs 220 2.75 60@1100@500 224
Na
2/3
[Mg
0.28
Mn
0.72
]O
2
SIBs 190 2.9 83@38@150 225
Na
0.6
[Cr
0.6
Ti
0.4
]O
2
SIBs 167 3.6 75@16.7@200 226
Na
0.80
Li
0.12
Ni
0.22
Mn
0.66
O
2
SIBs 214 3.4 115@21.4@200 227
Na
0.70
Mn
0.60
Ni
0.30
Co
0.10
O
2
SIBs 179 3.5 125@90@10 228
K
0.41
CoO
2
PIBs 103 0.56 3.0 56@18@30 229
K
0.45
MnO
2
PIBs 116 0.5 2.7 66@20@100 230
K
0.28
MnO
2
0.15H
2
O PIBs 77 2.5 71.5@20@100
KCrO
2
PIBs 218 0.59 2.5 60@10@100 231
K
0.5
V
2
O
5
PIBs 67 0.6 2.7 70@20@80 232
K
0.45
Mn
0.5
Co
0.5
O
2
PIBs 114 2.3 47@500@500 233
K
0.65
Fe
0.5
Mn
0.5
O
2
PIBs 154 2.2 55@100@350 234
K
0.44
Ni
0.22
Mn
0.78
O
2
PIBs 112 2.4 52@200@500 235
K
0.67
MnO
2
PIBs 237 0.61 2.6 78@50@300 236
K
1.39
Mn
3
O
6
PIBs 118 2.6 80@50@50 237
K
0.67
MnO
2
PIBs 159 0.64 2.5 40@200@400 238
K
0.5
MnO
2
PIBs 125 0.64 2.5 40@200@400
KCrO
2
PIBs 218 2.5 60@10@100 231
K
0.77
MnO
2
PIBs 176
K
0.32
MnO
2
PIBs 86
K
0.6
CoO
2
PIBs 141 2.5 65@100@300 232
KCrS
2
PIBs 173 2.4 60.5@35@300 239
K
0.69
CrO
2
PIBs 167 0.68 2.6 74@100@300 240
K
0.8
CrO
2
PIBs 186 2.5 50@186@300 241
KVO PIBs 253 0.52 2.5 87@10 mA g@50 242
K
0.5
V
2
O
5
PIBs 67 1.0 3.2 79.2@100 mA g@100 243
K
2
V
3
O
8
PIBs 149 0.53 1.75 60@20 mA g@200 244
K
0.83
V
2
O
5
PIBs 104 2.5 54@100 mA g@200 245
K
0.48
Mn
0.4
Co
0.6
O
2
PIBs 119 0.5 2.5 27@12@180 246
K
0.45
Mn
0.9
Mg
0.1
O
2
PIBs 119 2.4 20@20@100 247
K
0.7
Mn
0.7
Mg
0.3
O
2
PIBs 179 2.75 83.4@100@400 248
K
2/3
[Ni
1/3
Mn
2/3
]O
2
PIBs 155 3.0 74@85@200 249
K
0.83
[Ni
0.05
Mn
0.95
]O
2
PIBs 187 0.64 3.0 120@52@200 250
K
0.67
Mn
0.83
Ni
0.17
O
2
PIBs 158 2.1 52@500@200 251
K
0.45
Mn
0.5
Co
0.5
O
2
PIBs 113 2.3 82@50@50 252
K
0.67
Ni
0.17
Co
0.17
Mn
0.66
O
2
PIBs 157 2.9 76@20@100 253
LiFePO
4
-with Na doping LIBs 170 1.0 3.45 136@1700@3000 254
LiFePO
4
-with Ni doping LIBs 170 1.0 3.4 150@1700@5000 255
Cl-doped LiFePO
4
LIBs 170 1.2 105@1700@500 256
Nb-doping LIBs 170 3.4 114.3@1700@300 257
LiMnPO
4
LIBs 170 0.52 4.1 80@170@500 258
LiMn
1x
Fe
x
PO
4
LIBs 171 0.52 3.8 160@17@50 259
Li
0.92
Co
0.8
Fe
0.2
PO
4
LIBs 154.1 0.52 4.8 121@15@500 260
LiCoPO
4
LIBs 167.5 0.51 4.8 60@84@140 261
LiNiPO
4
LIBs 167.5 0.55 262
LiNi
0.5
Co
0.5
PO
4
LIBs 160
Li
3
V
2
(PO
4
)
3
LIBs 197 0.51 3.9 57@133@1000 263
Li
2
MnP
2
O
7
LIBs 110 0.44 4.7 (T) 264
Li
2
NiPO
4
F LIBs 143 0.44 265
Li
2
CoPO
4
F LIBs 143 0.44 266
LiCuSO
4
F LIBs 145 0.50 5.1 (T) 267
LiCoSO
4
F LIBs 149 0.56 4.7 (T) 268
LiNiSO
4
F LIBs 149 0.54 5.2 (T)—
LiNiOSO
4
LIBs 152 5.0 (T) 269
LiCoOSO
4
LIBs 152 5.1 (T)—
Li
2
NiOSO
4
LIBs 152 5.0 (T)—
LiNi
0.5
Mn
1.5
O
4
LIBs 147 4.7 110@140@200 270
LiNiPO
4
LIBs 167.5 0.55 5.1 110@167@100 271
LiNi
0.5
Co
0.5
PO
4
LIBs 167.5 0.55
Li
2
CoPO
4
F LIBs 287 0.47 4.8 50@2870@50 272
LiNiSO
4
F LIBs 148 0.54
Li
3
V
2
(PO
4
)
3
LIBs 132 0.62 4 90@66@40 273
Review Article Chem Soc Rev
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
This journal is © The Royal Society of Chemistry 2024 Chem. Soc. Rev.
Table 1 (continued )
Materials Battery types
Theoretical
capacity
(mA h g
1
)SNPC
size (nm) Average
potential (V)
Capacity (mA h g
1
)@
current density (mA g
1
)@
cycle number Ref.
Na-deficient LIBs 117 3.0 65@110@1000 274
Na
3.32
Fe
2.11
Ca
0.23
(P
2
O
7
)
2
Na deficient LIBs 194 2.8 106@550@300 275
Na
3.41
d
0.59
FeV(PO
4
)
3
NaFePO
4
SIBs 154 2.7 110@30@50 276
NaFePO
4
SIBs 154 0.38 2.3 140@20@300 277
a-FePO
4
SIBs 178 2.3 54@50@300 278
VOPO
4
SIBs 117 0.6 3.5 100@16.5@250 279
NaVOPO
4
SIBs 145 0.51 3.5 75@72@1000 280
Na
3
V(PO
4
)
2
SIBs 173 0.69 3.25 70@173200 281
Na
3
V
2
(PO
4
)
3
SIBs 117 0.52 3.25 48@118@5000 282
Na
3
V
2
(PO
4
)
3
SIBs 117 0.54 3.3 30@4680@30000
Na
3
Fe
2
(PO
4
)
3
SIBs 115 0.58 2.8 96@115@200 283
Na
3
Cr
2
(PO
4
)
3
SIBs 117 0.63 4.5 5@60@20 284
Na
2
TiV(PO
4
)
3
SIBs 178 2.2 73@1250@500 285
Na
3
FeV(PO
4
)
3
SIBs 111 2.6 100@110@1000 286
Na
4
MnV(PO
4
)
3
SIBs 111 3.4 90@110@1000 286
Na
3
MnZr(PO
4
)
3
SIBs 107 0.64 3.5 92@50@500 287
Na
3
MnTi(PO4)
3
SIBs 177 0.63 2.4 82@3500@3500 288
Na
2
MnP
2
O
7
SIBs 195 0.62 3.0 90@39@30 289
Na
2
CoP
2
O
7
SIBs 96.1 0.51 2.6 85@5@10 290
Na
3.12
Fe
2.44
(P
2
O
7
)
2
SIBs 110 3.0 92@220@80 291
Na
7
V
3
(P
2
O
7
)
4
SIBs 80 4.0 52.5@80@600 292
t-Na
2
(VO)P
2
O
7
SIBs 93.4 0.5 3.5 66@4.6@10 293
NaVP
2
O
7
SIBs 108 3.7 30@5.4@20 294
Na
4
Co
3
(PO
4
)
2
(P
2
O
7
) SIBs 127 4.5 70@12.7@100 295
Na
4
Ni
3
(PO
4
)
2
(P
2
O
7
) SIBs 129 4.6 100@10@10 296
Na
7
V
4
(P
2
O
7
)
4
PO
4
SIBs 92.8 3.8 81.4@46@300 297
NaVPO
4
F SIBs 143 0.55 3.5 101@286@1000 298
Na
3
V
2
O
2
(PO
4
)
2
F SIBs 130 3.7 102@2600@2000 299
Na
2
FePO
4
F SIBs 124 3.25 300
Na
2
CoPO
4
F SIBs 122 0.59 4.3 40@61@20 301
Na
2
Fe
2
(SO
4
)
3
SIBs 118 3.5 80@12@50 302
Na
2
Fe(SO
4
)22H
2
O SIBs 118 3.2 61@12@20 303
Na
2
Fe
2
(SO
4
)
3
SIBs 120 3.7 97@24@300 304
Na
2.5
Fe
1.75
(SO
4
)
3
SIBs 106 3.6 90@106@200 305
Na
6
Fe
5
(SO
4
)
8
SIBs 120 3.7 87.1@240@1000 306
NaFe(SO
4
)
2
SIBs 99 0.64 3.0 78@20@80 307
Na
2
MnSiO
4
SIBs 278 3.0 140@278@500 308
Na
2
FeSiO
4
SIBs 276 2.4 88@70@200 309
Na
2
CoSiO
4
SIBs 272 0.47 3.3 105@5@25 310
Na
4
Fe
3
(PO
4
)
2
(P
2
O
7
) SIBs 129 3.1 65@2600@4400 311
Na
2
Fe
2
(SO
4
)
3
SIBs 120 3.5 70@60@100 302
KTi
2
(PO4)
3
PIBs 64 0.62 1.7 80@32@100 312
KFeSO
4
F PIBs 255 0.5 3.6 52@20@100 313
K
3
V
2
(PO
4
)
3
PIBs 106 0.47 3.6 50@20@100 314
K
3
V
2
(PO
4
)
2
F
3
PIBs 115 0.58 3.75 59@10@50 315
KVOPO
4
PIBs 133 0.55 3.7 104@66@100 316
Fe
4
[Fe(CN)
6
]
3
5.89H
2
O LIBs 125 1.03 3.0 68@100@10 317
Fe
4
[Fe(CN)
6
]
3
LIBs 125 1.02 3.0 70@25@94 318
Fe[Fe(CN)
6
]
0.71
LIBs 125 1.02 3.0 20@25@30 319
Fe[Fe(CN)
6
]
0.87
0.133.1H
2
O LIBs 125 1.02 3.0 96@25@50
LiFeHCF-1 LIBs 170 1.02 3.0 109@190@650 320
LiFeHCF-2 LIBs 170 1.02 3.0 87@190@650
LiFeHCF-3 LIBs 170 1.02 3.0 41@190@650
LiFeHCF-4 LIBs 170 1.02 3.0 61@190@650
LiFeHCF-5 LIBs 170 1.02 3.0 29@190@650
NiFe-PBA SIBs 70 1.03 3.1 100@20@200 321
PB-1 SIBs 112 2.9 70@100@100 315
Na
2
Zn
3
[Fe
II
(CN)
6
]
2
9H
2
O SIBs 64.8 3.2 96@10@50 322
Ni-based PBAs SIBs 85 3.4 72@85@4000 323
PB-S1 SIBs 170 3.1 70@100@200 324
PB-S3 SIBs 170 2.9 72@100@500
Fe-based PBAs SIBs 157 3.0 80@150@500 325
Mn-based PBAs SIBs 140 1.06 3.3 74@500@2700 326
Na
1.7
FeFe(CN)
6
SIBs 148 1.06 2.9 99@200@100 327
Na
1.54
FeFe(CN)
6
SIBs 170 1.06 3.0 84@200@500 328
Na
2
Fe
II
[Fe
II
(CN)
6
] SIBs 170 2.9 71@50@100 329
Chem Soc Rev Review Article
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2024
Table 1 (continued )
Materials Battery types
Theoretical
capacity
(mA h g
1
)SNPC
size (nm) Average
potential (V)
Capacity (mA h g
1
)@
current density (mA g
1
)@
cycle number Ref.
NaFe
III
[Fe
II
(CN)
6
] SIBs 140 1.0 3.0 50@50@100
Fe
III
[Fe
III
(CN)
6
] SIBs 64.8 1.0 3.2 6@50@100
PB PIBs 87 3.6 80@9@500 330
K
1.64
Fe[Fe
II
(CN)
6
]
0.89
0.15H
2
O PIBs 155 1.0 3.3 130@30@100 331
K
1.75
Mn[Fe
II
(CN)
6
]
0.93
0.16H
2
O PIBs 155 1.0 3.6 120@30@100
K
2
Mn[Fe(CN)
6
] PIBs 155 1.02 3.4 150@15@140 332
KFe
III
Fe
II
(CN)
6
PIBs 155 1.02 3.3 120@10@100 333
KFe[Fe(CN)
6
] PIBs 155 1.02 3.3 30@500@1100 334
K
0.6
Ni
1.2
[Fe(CN)
6
] PIBs 155 3.3 62@50@300 335
RGO@PB PIBs 87 3.3 62@50@300 335
K
1.75
Mn[Fe(CN
6
)]0.930.16H2O PIBs 155 1.02 3.8 128@30@100 331
K
1.64
Fe[Fe(CN)
6
]
0.89
0.15H
2
O PIBs 155 1.02 3.3 110@30@100 331
KMHCF PIBs 156 3.7 85@156@100 336
NI-KMHCF PIBs 156 3.7 102@156@100
K
1.6
Mn [Fe(CN)
6
]
0.96
0.27H
2
O PIBs 131 3.9 83@50@30 337
K
0.220
Fe[Fe(CN)
6
]
0.805
4.01H
2
O PIBs 125 0.51 3.2 117@125@100 338
FeFe-PW PIBs 119.7 3.3 90.4@20@100 339
CoFe-PW PIBs 108.2 3.4 43@20@15
NiFe-PW PIBs 70.7 3.7 64@20@15
CuFe-PW PIBs 59.1 3.7 30@20@15
KFeHCF-S PIBs 78@100@300 340
KFeHCF-M PIBs 108@20@15
KFeHCF-L PIBs 8@20@15
PTCDA LIBs 273 0.35 2.4 120@100@280 341
IEP-11-E12 LIBs 1–2 2.2 46.7@550@9000 342
CMP LIBs 1.69 101@1000@1500 343
Py-A-CMP LIBs 2.1 2.2 125@100@400 344
TPE-A-CMP LIBs 1.56 2.2 170@100@400
PTTPAB LIBs 0.75B1 3.8 95@20@50 345
SPTPA LIBs 1.0 3.75 88@50@1700 346
YPTP LIBs 109.4 1.1, 1.6 3.75 92@50@1700
OPTPA LIBs 109.4 1.7 3.75
SPTPA LIBs 109.4 1, 1.7 3.75
POP SIBs 1.05 1 180@30@150 347
POP SIBs 1.4–2 2 200@10@40 348
TAPT-NTCDA SIBs 1.25 1.5 70@50@50 349
PI PIBs 4.7 2.2 48.7@50@200 350
PQI PIBs 4.7 2.2 84.7@50@200
PI-CMP PIBs 4.7 2.25 93.3@50@200
HAT PIBs 245 0.7–1.4 1.7 169@10000@4600 351
PIBN-G (COF) LIBs 280 1.4 2.3 242.3@280@300 352
BQ1-COF LIBs 773 1.2 2.2 230@1556@1000 353
PIBN LIBs 280 1.45 2.0 206@280@300 352
CT SIBs 454 1.3 2.25 140@500@100 354
N-COF SIBs 515 1.1 1.5 236.5@500@1000 355
DAAQ-COF@CNT PIBs 2.3 1.5 108@500@500 356
COF PIBs 1.95 1.5 70@1000@5000 357
MIL-53(Fe) LIBs 0.65 2.6 75@75@50 358
MIL-132 LIBs 59 3.0 50@10@5 359
MOF-1 LIBs 2 2.3 40@100@50 360
MOF-2 LIBs 2 2.4 37@100@50
MOF-3 LIBs 2 2.5 8@100@50
MOF-4 LIBs 2 2.2 5@100@50
S/MC LSBs 1675 2 1.4 500@167@800 361
AB-S LSBs 1675 2 1.8 400@837@50 362
MICP LSBs 678 1.1 1.8 341@67.8@3000 363
S/OBC LSBs 1675 2.3 2.0 711@167@50 364
S/TBC LSBs 1675 1.4 1.9 401@167@50
S/DBC LSBs 1675 2.5 2.1 291@167@50
FDU/S-40 LSBs 1675 0.46 1.6 780@400@50 365
MPC/Se LSBs 678 1–3 1.7 530@67.8@100 366
ACC SSBs 1675 0.5 1.1 700@1675@2000 367
S/ELSC-40 SSBs 1675 0.7 1.0 518@1675@500 368
APCF-38S SSBs 1675 0.3–0.8 0.9 1000@167@400 369
WBMC SSBs 1675 0.4–1.0 0.9 822@334@100 370
C/S PSBs 1675 0.4–1.0 0.6 868@20@150 371
Review Article Chem Soc Rev
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
This journal is © The Royal Society of Chemistry 2024 Chem. Soc. Rev.
Table 2 Summary of the anode materials with SNPCs
Materials Battery types
Theoretical
capacity
(mA h g
1
)SNPC
size (nm) Average
potential (V)
Capacity (mA h g
1
)@
current density (mA g
1
)@
cycle number Ref.
TiS
2
LIBs 956 0.57 1.6 213@100@200 372
ZnSe/CoSe
2
-CN SIBs 433 0.38/0.58 1.0 422@500@300 373
TiO
2
SIBs 335 0.3 165@6700@2000 374
CoSi
3
P
3
LIBs 2902 0.34 0.3 891@100@200 375
Co
3
O
4
LIBs 890 0.8 376
TiNb
2
O
7
LIBs 387.6 2 1.6 377
BP LIBs 2596 1 0.8 600@1000@100 378
MoO
2
LIBs 838 1 0.4 489@100@1050 379
MoS
2
SIBs 670 0.64 0.3 172@200@200 380
MoS
2
LIBs 670 0.69 1.6 205@200@1400 381
MoSe
2
LIBs 422 0.64 0.6 494@422@100 382
e-MoSe
2
0.98 0.8 80@422@100
MoSe
2
@C@MoO
x
SIBs 422 0.65 0.6 450@200@200 383
MoSe
2
@C SIBs 422 0.64 0.6 200@200@200
MoSe
2
SIBs 422 0.64 0.6 320@200@150
Ti
2
AlC
3
SIBs 351 0.93 0.5 384
Ti
3
C
2
SIBs 351 0.99 0.5
a-Ti
3
C
2
SIBs 351 1.25 0.5 50@200@500
Ti
3
C
2
PIBs 191 0.99
a-Ti
3
C
2
PIBs 191 1.25 0.25 47@200@500
Mo
2
TiC
2
PIBs 356 0.99 0.5 140@20@94 385
Pillared-Mo
2
TiC
2
PIBs 356 1.12 0.4 250@20@94
Ti
3
C
2
T
x
SIBs 351 1.2/0.97 0.5 100@20@100 386
Ti
3
C
2
T
x
SIBs 351 0.99 0.4 76@100@120 387
Ti
3
C
2
T
x
PIBs 191 0.99 0.3 42@100@120
Ti
2
C SIBs 331 0.77/1.0 0.6 175@20@100 388
V
2
C SIBs 335 0.952 0.8 21@1000@300 389
Carbon SIBs 297 0.3/0.5 0.2 280@100@300 390
TiO
2
LIBs 335 0.95 391
Hard carbon LIBs 0.62/0.6 0.01 500@217@200 392
a-C-900 SIBs 0.55 0.6 280@100@5 393
a-C-1300 SIBs 0.55 0.7 270@100@5
a-C-1800 SIBs 0.55 0.7 280@100@5
Hard carbon PIBs 0.42 0.2 200@100@100 394
Graphite PIBs 0.34 0.25 16@100@100
SiC-CDC-800 PIBs 0.52 0.5 284@100@200 395
SiC-CDC-900 PIBs 1.19 0.4 198@100@200
SiC-CDC-1000 PIBs 1.53 0.5 164@100@200
L-A-900 PIBs 0.4 0.3 118@1000@1200 396
B1-A-900 PIBs 0.5 0.3 95@1000@1200
B2-A-900 PIBs 0.45 0.4 52@1000@1200
Graphene SIBs 0.4 0.6 120@50@50 397
C-900 SIBs 0.6 0.5 150@50@50
C-1300 SIBs 0.7 0.1 297@50@50
Hard carbon SIBs 0.5 0.1 250@50@200 398
PDCzBT SIBs 1 0.2 117@100@200 399
LIBs 1 0.25 300@200@400
DBD-CMP1 LIBs 0.4–0.6 0.2 600@100@300 400
DBD-CMP2 LIBs 0.25 400@100@300
DBD-CMP1 SIBs 0.4–0.6 0.3 241@100@100
DBD-CMP2 SIBs 0.4 83@100@100
CMP-PyBT PIBs 0.94 0.75 104@500@500 401
CMP-PyBz PIBs 1.1 0.5 272@500@500
CTF-0 PIBs 0.5 0.3 113@100@200 402
CTF-1 PIBs 0.7 0.8 60@100@200
Si@ZIF-8 LIBs 4200 1.1 0.5 830@200@500 403
ZIF-8-C@PP LIBs 2596 0.8/1.3 0.25 786@100@100 404
P@NMC SIBs 2596 0.6/0.8 0.4 450@1000@1000 405
P@HC SIBs 2596 0.6/1.1 0.3 548@1000@1000 406
P@N-CNT LIBs 2595 0.9/1.1 0.3 523@5000@1500 407
OMC@RP SIBs 2596 0.75 0.4 600@5200@1000 408
Cu-OMC@RP SIBs 2596 1.2/1.4 0.5 750@5200@750
MnS LIBs 616 1–2 0.5 460@1600@1500 409
NiP
2
SIBs 1333 0.3 244@2000@1000 410
FeP LIBs 926 0.46 0.6 590@2000@1000 411
GeP
5
LIBs 2300 0.75 2200@200@40 412
ZnP
2
LIBs 1477 0.48 0.5 100@1000@200 413
CoNi-LDH LIBs 744 0.81 0.8 440@5000@500 414
Chem Soc Rev Review Article
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2024
layered oxide cathodes, optimizing the size of the SNPC
between the adjacent MO
2
layers to increase the structure
stability and reaction kinetics is essential. Therefore, except
for the unary transition metal oxides, binary transition metal
oxides (i.e., Mn/Ni, Mn/Fe, and Ni/Fe-based oxides) and ternary
transition metal oxides (LiNi
x
Co
y
Mn
1xy
O
2
) were also inten-
sively investigated to adjust the SNPC for fast ion diffusion
kinetics, which increases both high electrochemical perfor-
mance and low-cost advantages.
556–558
Especially, the well-
recognized ternary transition metal oxides were successfully
applied in industry for large-scale battery packages.
559,560
The above SNPC designs alleviate the structure distortion
and damage to some degree during ion diffusion. When the Li
+
applies as the transport ion ‘A’, the diffusion of the ions inside
the MO
2
layers slowly distorts the order of the layered structure,
which is still affordable for rechargeable batteries. However, if
Na
+
and K
+
ions with much larger ionic radii are used as
transport cations, the distortion of MO
2
layers becomes severe,
damaging the SNPC for the ion transport and significantly
shortens the service life of the layered oxide cathodes. Further
development of layered oxide materials with appropriate SNPCs
to facilitate ion diffusion is required.
Currently, heteroatom doping is one of the most popular
strategies to improve the electrochemical performances of
layered oxide cathodes with SNPCs. Many elements, such as
Mg, Ti, Cu, Sn, Zn, etc., have been investigated to optimize the
interlayer distances between adjacent MO
2
layers to stabilize
the sub-nanochannels for ion diffusion. Su et al. demonstrated
that the doping of the Mg element into CoO
2
layers to form
the Na(Co
0.92
Mg
0.08
)O
2
cathode material could change the
electron density and the valence state of oxygen, and effectively
improve the Na
+
ion diffusion, making it seven times faster by
enlarging and stabilizing the SNPC with stronger Coulombic
repulsion between the adjacent oxide layers (Fig. 8b).
552
Hu
et al. designed the Ti-substituted layered oxide cathode
(Na
0.8
Li
0.27
Mn
0.68
Ti
0.05
O
2
) and indicated that the doping of
the Ti element significantly relieved the structure distortion
due to the Jahn–Teller effect, effectively stabilizing the SNPC at
around 0.55 nm for ion transport (Fig. 8c).
553
Furthermore,
Chen et al. discovered that doping Cu element into MO
2
layers
enlarged the SNPC, preventing the shrinkage of the SNPC
during the stripping of ions and increasing the mobility of
Na
+
ions. As a result, the Cu-substituted layered oxide cathode
(Na
0.67
Mn
0.6
Ni
0.2
Co
0.1
Cu
0.1
O
2
) provided a capacity retention of
80.0% after 500 cycles at a high current density of up to 1 A g
1
in SIBs.
562
Yun et al. indicated that the doping of Sn into
lithium-rich layered oxide (Li
2
IrO
3
) effectively enlarged the
SNPC and prevented structural collapse at high charging vol-
tage, which benefited from the migration of Sn between the
adjacent IrO
2
layers during cycling (Fig. 8d).
554
Moreover, Zheng
et al. reported that Zn-substitution in MO
2
layers of materials
(K
0.02
Na
0.55
Mn
0.70
Ni
0.25
Zn
0.05
O
2
) reduced the anisotropic coupling
between Mn
4+
and oxidized O
2
. This Zn doping strategy effec-
tively stabilized the SNPC and provided the layered oxides with a
superior cycling performance in K-ion batteries (97% capacity
retention after 1000 cycles at 100 mA g
1
).
563
Besides the heteroatom doping strategy, an innovative
method to stabilize SNPCs for ion diffusion is to further modify
the MO
2
layers through high-entropy design. The high-entropy
design achieved by doping five or more elements to form a
single phase and achieve a high entropy situation in layered
oxides aims to provide super stable MO
2
layers with homo-
genous internal stress for fulfilling the requirement of con-
structing effective SNPCs for fast ion transport. For example,
Zhao et al. reported a high entropy sodium layered oxide
(NaNi
0.12
Cu
0.12
Mg
0.12
Fe
0.15
Co
0.15
Mn
0.1
Ti
0.1
Sn
0.1
Sb
0.04
O
2
)by
mixing nine elements (Ni, Cu, Mg, Fe, Co, Mn, Ti, Sn, and
Sb) into MO
2
layers, which forms very stable SNPCs for fast Na
+
ion diffusion.
561
The high-entropy MO
2
layers effectively
delayed the phase transformation from O3 to P3 during cycling
and maintained around 60% capacity in the O3 phase region
(Fig. 8e and f). Consequently, the high-entropy layered oxide
Table 2 (continued )
Materials Battery types
Theoretical
capacity
(mA h g
1
)SNPC
size (nm) Average
potential (V)
Capacity (mA h g
1
)@
current density (mA g
1
)@
cycle number Ref.
CoFe-LDH SIBs 0.7 0.8 209@1000@200 415
CoFe-LDH LIBs 0.92 0.8 860@4000@1000 416
Graphene sheets LIBs 0.57 0.3 718@500@40 417
Graphene sheets SIBs 0.37 0.25 917@50@100 418
Graphene sheets PIBs 0.35 0.35 400@500@600 419
Graphite LIBs 372 0.335 0.1 306@1000@800 420
Natural graphite PIBs 278 1.2 0.7/0.75 90@500@100 414
Graphite SIBs 278 0.335 1.1/1.2 125@100@100 421
Graphite SIBs 0.37 0.3 300@20@30 422
MoP
2
LIBs 676 0.56 0.6 525@160@60 423
FeP LIBs 0.7 750@200@100 424
SnP
0.94
LIBs 488 0.4 0.4 670@120@40 425
NaTi
2
(PO
4
)
3
SIBs 485 0.366 426
Na
2
Ti
3
O
7
SIBs 311 0.845 427
Li
4
Ti
5
O
12
LIBs 175 0.835 0.75 220@125@500 428
MnO
2
SIBs 308 0.693 1.25 72.2@500@100 357
F-COF LIBs 314 1.93 0.2 95@1000@5000 357
COF LIBs 421 1.95 0.2 70@1000@4000 357
Review Article Chem Soc Rev
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
This journal is © The Royal Society of Chemistry 2024 Chem. Soc. Rev.
Table 3 Summary of the electrolytes with SNPCs
Electrolyte SNPC size (nm) Ion types Ionic conductivity (S cm
1
) Ref.
PEO-LiTFSI 0.7 Li
+
429
TEA-TFB 0.6 TEA
+
430
1.1 BF
4
SiN
x
film 0.45 H
+
431
Na
+
V-CNF 0.6 Zn
2+
0.61 10
3
432
1.3
Zeolite 0.54 Li
+
0.94 10
3
433
UiO-66-X 0.6 F
,Cl
B10 10
3
434
CuBTC MOF 0.65 Li
+
B5.0 10
4
435
MOF/polymer 0.8 Li
+
0.54 10
3
18
PvDF-LiTFSI-MOF 1.16 Li
+
4.08 10
4
436
Ce-UiO-66-Li
+
0.6 Li
+
2.16 10
4
437
PEO: LiAsF
6
0.5 Li
+
438
Cellulose nanofibrils 0.9 Li
+
1.5 10
3
439
MOF-cellulose nanofibers 1.1 Li
+
7.88 10
4
440
PP/UiO-66-NH
2
0.4 Li
+
2.45 10
4
441
Li
9.54
Si
1.74
P
1.44
S
11.7
Cl
0.3
0.8 Li
+
2.5 10
3
442
Li
3
OCl 0.4 Li
+
6.6 10
5
443
Li
1+x
Ti
2x
Al
x
(PO
4
)
3
0.5 Li
+
444
POSS-PEO/LiTFSI 0.74 Li
+
5.6 10
6
445
POSS-IL-LiTFSI Li
+
4.8 10
4
445
EMIM-Cl@UiO-67 0.65 Li
+
0.68 10
4
446
F-IL-GEL Li
+
9.16 10
3
447
PIM-EA-TB FSI
0.5 10
2
448
Li
1.5
Al
0.5
Ge
1.5
P
3
O
12
0.6 Li
+
410
4
449
GF@ZIF@PEO 0.38/1.31 Li
+
1.78 10
4
450
GF@PEO 0.74 Li
+
1.60 10
4
450
PEO-E 0.74 Li
+
2.03 10
4
450
MOF-LiCl 2 Li
+
2.4 10
5
451
MOF-LiBr Li
+
3.2 10
5
451
MOF-LiI Li
+
1.1 10
4
451
PVDF-HFP/LLZO/LiTFSI 3.41 Li
+
4.9 10
4
452
Mg
2
(dobdc) 1.1/1.3 Li
+
3.1 10
4
453
MOF 1.0/2.0 Li
+
7.41 10
4
454
Li–IL@UIO-67 1.2/1.6 Li
+
4.3 10
4
455
EMI-TFSA 1.16 Li
+
0.5 10
4
456
Li-IL@MOF 1.2 Li
+
3.0 10
4
457
SLE-H 0.6 Li
+
3.3 10
4
458
SIL/UIO-66 QSSE 0.6 Li
+
3.7 10
4
459
LIM-L 1.2 Li
+
1.0 10
4
460
MIL-121/Li 0.87 Li
+
510
4
461
Li
6.75
La
3
Zr
1.75
Ta
0.25
O
12
—Li
+
1.2 10
4
462
LLZO-polymer 1.2 Li
+
10
4
–10
3
463
PEO 0.74 Li
+
10
5
464
PAN 1.1 Li
+
1.07 10
5
465
PVDF 0.56 Li
+
7.27 10
4
466
PAN Li
+
6.5 10
4
460
PEO 0.74 Li
+
1.4 10
3
467
PEGDA Li
+
2.26 10
4
462
NaPF
6
-PEO 0.74 Na
+
510
6
468
NaTFSI-PEO 0.74 Na
+
4.5 10
6
469
NaFNFSI-PEO 0.74 Na
+
210
6
470
NaTCP-PEO 0.74 Na
+
6.9 10
5
471
NaClO
4
-PVP 1.2/1.3 Na
+
2.5 10
6
472
NaBr-PVA 0.34 Na
+
1.36 10
6
473
Na
3
Zr
2
Si
2
PO
12
0.37 Na
+
1.2 10
3
474
Na
3.1
Zr
1.95
Mg
0.05
Si
2
PO
12
0.65 Na
+
3.5 10
3
475
Na
3.4
Zr
1.6
Sc
0.4
(SiO
4
)
2
(PO
4
) 0.54 Na
+
410
3
476
Na
3
PS
4
0.4/0.5 Na
+
3.9 10
4
477
94Na
3
PS
4
6Na
4
SiS
4
0.5 Na
+
7.4 10
4
478
Na
3
P
0.62
As
0.38
S
4
0.7 Na
+
1.46 10
3
479
Na
2.9375
PS
3.9375
Cl
0.0625
0.4/0.5 Na
+
1.14 10
3
480
Na
3
Pse
4
0.73 Na
+
1.16 10
3
481
Na
2
B
10
H
10
0.98 Na
+
110
2
482
Na
2
(B
12
H
12
)
0.5
(B
10
H
10
)
0.5
0.58 Na
+
910
4
483
PVDF-HFP 0.3 Na
+
110
3
477
PEO/Na
3.4
Zr
1.8
Mg
0.2
Si
2
PO
12
0.74/0.4 Na
+
2.4 10
3
484
PEO/KbrO
3
0.74 K
+
7.74 10
8
485
PEO/KBr 0.74 K
+
5.0 10
7
486
Chem Soc Rev Review Article
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2024
cathode delivered a high-capacity retention of around 83% after
500 cycles at 3C-rate in SIBs.
3.1.2. Tuning SNPCs in polyanionic and spinel-like cath-
odes. Polyanionic materials are another class of cathode mate-
rials coupled with SNPCs to realize redox reactions and cation
storage in rechargeable batteries, showing different advantages
compared to layered oxides. Applying polyanionic materials as
cathodes has attracted great interest since the late 1990s, due to
their high operating potential, stable SNPCs (i.e., 3D open
framework) with less volume change and no phase transition,
high thermal stability, and high safety.
564–566
However, the
drawbacks of polyanionic materials are their low specific
capacities, low electronic conductivities, and poor low-
temperature performances, which restrict their wide applica-
tion under harsh environmental conditions.
Polyanionic materials A
x
M
y
(XO
4
)
n
consist of tetrahedral
polyanionic units (XO
4
)
n
and covalently bonded polyhedra
MO
6
(where A represents Li, Na, and K, X represents P, S, Si,
etc., and M represents transition metals) with SNPCs of around
0.4 to 0.6 nm (Fig. 7b).
311,567–571
This special structure of
polyanionic materials leads to low conductivity because of
the electronic interactions between (XO
4
)
n
units during the
charge/discharge process. According to X elements, the poly-
anionic materials can be classified as phosphate with X = P
(orthophosphates, fluorophosphates, pyrophosphates, and
mixed pyrophosphates), sulfates with X = S (AM(SO
4
)
2
,
A
2
M
2
(SO
4
)
3
), and silicates with X = Si (A
2
MSiO
4
), which share
similar SNPC structures.
566
Similar to the development of
layered metal oxides, the heteroatom doping strategy is
the most popular way for improving the structure stability
and ionic conductivity of the SNPCs inside polyanionic cath-
odes. However, to further improve the electronic and ionic
conductivity of the polyanionic materials, doping modification
is commonly accompanied by carbon coating and particle size
control strategies.
The heteroatom doping strategy used in polyanionic cathode
materials can be categorized as (1) doping in transport ion sites
and (2) doping in transition metal sites. The doping in trans-
port ion sites involves incorporating heteroatom ions with large
radii and high conductivity to control the size of SNPCs and act
as the ‘pillars’ to stabilize the ion diffusion paths. For instance,
Lim et al. doped the K element into the Na
3
V
2
(PO
4
)
3
/C (NVP/C)
cathode material at the transport Na
+
ion sites to form K
0.09
-
NVP/C.
567
Due to the larger ionic radius of K than that of Na,
the K elements act as the pillar and enlarge the SNPC for Na
diffusion and accommodation (Fig. 9a). With the pillar effect
provided by K doping, the doped cathode material showed a
significantly smaller volume change than that of the pristine
materials (reduced from 9.02% to 6.64%) (Fig. 9b). Conse-
quently, the K
0.09
-NVP/C cathode achieved a high reversible
capacity of 83 mA h g
1
after 200 cycles at the 1C-rate.
Doping heteroatom ions in transition metal sites aims to
stabilize the framework structure and increase the conductivity
of the SNPC. For example, Shen et al. designed a NASICON-type
cathode (Na
3
V
2x
Mn
x
(PO
4
)
3
/C) by doping Mn
2+
ions into the
transition metal sites of Na
3
V
2
(PO
4
)
3
/C (Fig. 9c).
568
The doping
Table 3 (continued )
Electrolyte SNPC size (nm) Ion types Ionic conductivity (S cm
1
) Ref.
PEO/CH
3
COOK 0.74 K
+
2.74 10
7
487
PVP/KIO
3
—K
+
110
9
488
PPCB/KFSI 0.74 K
+
1.36 10
5
489
PEO/KNO
3
/KI 0.74 K
+
6.15 10
6
490
K
2
Fe
4
O
7
0.4 K
+
3.5 10
2
491
PAF-220 1.3/1.9 Li
+
210
4
492
PAF-220-Li 1.9 Li
+
510
4
492
PEO-LiTFSI 0.74 Li
+
2.3 10
4
493
Halide 0.452 Li
+
1.0 10
4
493
Molecular sieve 0.32 Na
+
494
AO-PIM-1 0.45 K
+
,Na
+
,Li
+
,Cl
4.4 10
4
495
PVDF-HFP-CPT-[PMPyr][TFSI] Li
+
4.2 10
5
496
SSZ-13 + PEO + LiTFSI 0.74 Li
+
5.34 10
2
497
PEO + LiTFSI + ZYNa 0.74 Li
+
1.66 10
2
498
PHS Li
+
1.36 10
5
499
Mn-PBA 1.0 Na
+
9.1 10
5
500
Cubic Na
3
PS
4
0.4 Na
+
210
4
501
0.5
Na
2.4
Er
0.4
Zr
0.6
Cl
6
0.4 Na
+
3.5 10
5
502
Na
2.9
Sb
0.9
W
0.1
S
4
0.7 Na
+
4.1 10
2
503
Na
3x
Y
1x
Zr
x
Cl
6
0.2 Na
+
6.6 10
5
504
K
2.2
Ba
0.4
SbSe
4
0.5 K
+
4.45 10
5
505
Na
2
MgZnTeO
6
0.6 Na
+
1.4 10
5
506
Na
2
Mg
2
TeO
6
0.55 Na
+
2.3 10
4
507
CPCSE 0.95 Li
+
5.2 10
4
508
COF 0.9 Li
+
1.71 10
4
509
MOF 0.34 Li
+
6.7 10
4
510
Carbon 0.7 Li
+
159
PPS@TiO
2
0.3/0.4 Li
+
,Na
+
0.6 to 3 10
4
61
MLM 0.56 K
+
511
MLM-EDTA 0.6 K
+
511
Review Article Chem Soc Rev
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
This journal is © The Royal Society of Chemistry 2024 Chem. Soc. Rev.
of Mn
2+
effectively enlarged the SNPC (Fig. 9d), stabilized the
crystal structure, and promoted the ionic and electronic con-
ductivity due to the large radius of Mn
2+
(0.91 Å) compared to
that of V
3+
(0.64 Å) and self-polarization ability. As a result, the
Mn-doped cathode materials achieved a high reversible dis-
charge capacity of 92.5 mA h g
1
after 100 cycles at 5C, which is
around 95.9% capacity retention.
Besides heteroatom doping, another pathway to enlarge the
ion diffusion channels and increase the ion conductivity of
SNPC inside the polyanionic materials is by combining them
with high-conductive materials. Li et al. introduced carbon
nanotubes into Na
3
V
2
(PO
4
)
2
F
3
(NVPF), which significantly
improved the NVPF crystallinity and ion transport (Fig. 9e
and f).
569
The NVPF@CNT cathode offered a super cycling
performance with 77.3% capacity retention after 5000 cycles
at 20 C. Guo et al. designed a heterogeneous NASICON-type
composite by combining Na
3
V
2
(PO
4
)
3
with Na
3
Fe
2
(PO
4
)(P
2
O
7
)
(NVFPP).
570
The introduced NVFPP phase provided enlarged
SNPCs. As a result, after 500 cycles, their cathode delivered a
78.6% capacity retention rate at 5C.
Moreover, the conventional polyanionic cathode usually
contains the V element, which is expensive and bio-toxic.
Substituting the V element with other cheap and eco-friendly
transition metals while maintaining the stability and conduc-
tivity of SNPCs in polyanionic materials is also worth
studying for practical applications. Chen et al. reported a
Na
3.32
Fe
2.34
(P
2
O
7
)
2
/C composite as the cathode material with
obvious eco-friendly advantages. Their V-free cathode also
provided sufficient SNPCs for Na
+
ion transport.
571
Later, they
further designed an eco-friendly cathode with a tenable nano-
sized Na
4
Fe
3
(PO
4
)
2
(P
2
O
7
)/C (NFPP-E) composition, exhibiting
enhanced air stability and suitability for all-climate operation
conditions. After 4400 cycles at 20C, the SNPC of NFPPE could
be maintained at around 0.5 nm.
311
The robust 3D frameworks provide the polyanionic cathode
with a more stable SNPC than that of the conventional layered
oxide materials. In addition, the spinel engineering design can
provide a 3D framework for transition metal oxides and form
AM[M
2
]O
4
(M = Fe or Mn) spinel-like cathode materials, which
is worth mentioning here. In spinel-like materials, the [M
2
]O
4
units form a spinel framework that stabilize the whole struc-
ture, allowing the reversible alkali ion insertion and extraction.
The spinel-like cathodes cannot offer a high energy density as
the traditional layered oxide materials, but they can provide
relatively lower costs and high safety like the polyanionic
cathode.
574
Spinel-like cathodes with 3D SNPCs have been
commercialized in LIBs, and their application in sodium-ion
batteries is currently attracting researchers.
A popular strategy to improve the electrochemical perfor-
mance and SNPC stability of spinel-like cathodes is the
chemical substitution to form the A[Mn
2x
M
x
]O
4
, where M is
the substituted metal. For example, the well-known
LiNi
0.5
Mn
1.5
O
4
(LNMO) is Ni-substitution modified LiMnO
4
spinel-like cathodes. The Ni substitution significantly improves
the cathode’s capacity (B130 mA h g
1
) and operating voltage
(B4.7 V vs. Li). Furthermore, due to the high voltage and less
involvement of the Mn
4+
redox reaction during cycling, the
Jahn–Teller distortion is effectively reduced in the LNMO
cathode, stabilizing the SNPC for ion transport.
574,575
Chiring
and Senguttuvan reported a spinel-NaMnSnO
4
cathode
(Na
0.877
MnSnO
4
) for SIBs with Sn substitution (Fig. 9g).
572
The Sn
4+
in the spinel framework effectively reduced the ion
diffusion barriers, suppressed the Jahn–Teller effect with no
Table 4 Summary of the functional interlayers/separators with SNPCs
Materials SNPC
size (nm) Ion
types Ionic conductivity
(S cm
1
) Ref.
ZIF-7@PCF 0.3 Li
+
2.87 10
6
/8.09 10
7
512
N,S-Mo
2
C/C-ACF 1.2 Li
+
1.04 10
3
513
N-Mo
2
C/C-ACF 1.4 Li
+
1.31 10
3
N-C-AMP 0.5/0.8 Li
+
1.28 10
3
Celgard Li
+
7.6 0.15 10
4
514
P(VDF-HFP) Li
+
2.6 0.04 10
4
515
Li-Nafion Li
+
2.1 0.04 10
5
Li-PFSD Li
+
1.2 0.02 10
4
Z-PE 2 Na
+
7.0 10
4
516
UiO-66 MMM 0.8 Li
+
0.67 10
3
147
LMA 1.1 Li
+
2.310
6
149
MOF-coated LMA 1.1 Li
+
3.3 10
3
MOF@GO 0.9 Li
+
3.8 10
4
517
MOF@GO 0.9 Li
+
710
5
S-PE Li
+
3.5 10
4
518
S-GE 8 10
4
S-CE/S-PE 3 10
3
S-CE/S-GE 8 10
3
LLZO 0.35 Li
+
1.03 10
3
LLZTO 0.4 Li
+
210
3
Azo-TbTh 0.36 Na
+
6.9 10
3
519
TA/Fe
III
MOF-PP 1.9 Li
+
1.1 10
3
520
Spim-SBF-0.53 0.55 Li
+
5.5 10
3
521
Spim-SBF-0.98 0.54 Li
+
1.2 10
2
Spim-SBF-1.4 0.54 Li
+
2.6 10
2
Spim-SBF-1.67 0.54 Li
+
6.5 10
2
Spim-SBF-1.86 0.52 Li
+
7.6 10
2
Spim-SBF 0.55 Li
+
2.8 10
2
TiO
2
/PP 0.38 Li
+
0.246 10
3
61
MIL-125(Ti)-PP/PE 0.77 Li
+
522
Cpim-1 0.55, 0.9 4 10
2
523
AO-PIM-1 0.55, 0.9 5 10
6
MGEs-S0 2 Li
+
9.77 10
5
524
MGEs-S1 2 1.80 10
4
MGEs-S2 2 5.37 10
4
MGEs-S3 2 6.76 10
4
PAN-C Li
+
1.04 10
3
525
oxy-PAN-60 Li
+
0.461 10
3
oxy-PAN-120 Li
+
0.459 10
3
PVDF membrane Na
+
7.38 10
4
526
PVDF coated PP Na
+
1.25 10
3
527
PEO-KFSI 0.74 K
+
2.7 10
4
528
Poly (PC) (PPC)-KFSI K
+
1.36 10
5
529
Al
2
O
3
—Na
+
6.3–6.8 10
3
530
NaAlO
2
—Na
+
5.5–6.5 10
3
MATEPP/MMA/TFMA 0.74, 0.88 Na
+
6.29 10
3
531
LLTO/Li
3
PO
4
/polymer 1.2 Li
+
5.1 10
4
532
2D-SPE Li
+
7.14 10
5
533
SHCPE 0.3 Li
+
810
5
534
P(MMA-AN-VAc) 0.57, 0.88, Li
+
1.54 10
3
535
P(AN-co-MA) 1.1 Li
+
6.7 10
4
536
PVN-GPE 0.78 2.6 10
4
537
PDMS/PVDF 0.56 1.17 10
3
538
PEO/LiBOB/LLZTO 0,74 Li
+
2.4 10
5
539
PEGBCDMA based PEM Li
+
810
4
540
HPEI-PGC-PCL Li
+
5.36 10
4
541
PEO K
+
4.3 10
3
ZIF-67@PP 1.1 Li
+
1.64 10
3
542
Chem Soc Rev Review Article
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2024
observed change in the axial bonds in Mn
3+
centers (Fig. 9h),
and thus promoted highly stable SNPCs (B0.5 nm).
Furthermore, the coating strategy is also commonly used on
the surface of spinel-like cathodes against transition metal
Fig. 6 (a) The summary of SPNC size and theoretical capacities for oxides, polyanions, PBAs, polymers, and (b) the summary of SPNC size and capacities
for COFs, and S/Se cathodes in rechargeable batteries. (c) The summary of SPNC size and theoretical capacities of oxides, phosphides, sulfides, and
polymers, and (d) the summary of SNPC size and capacities of MXenes, graphite-based materials, and hard carbons in rechargeable batteries. (e) The
summary of SPNC size and ionic conductivity of polymers and MOF and, (f) the summary of SPNC size and ionic conductivity of NASICONs, sulfides, and
hydrides for quasi-solid-state and solid-state electrolytes in rechargeable batteries. (g) The summary of SPNC size and ionic conductivity of polymers and
oxides, and the summary of SNPC size and ionic conductivity for interlayers/separators in rechargeable batteries. (Red, green, and blue represent Li, Na,
and K, respectively.)
Review Article Chem Soc Rev
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
This journal is © The Royal Society of Chemistry 2024 Chem. Soc. Rev.
dissolution, impedance rise, and capacity fade under harsh
operating conditions such as high temperatures. For instance,
Chu et al. coated AlF
3
onto the LNMO.
573
They found that the
1 wt% AlF
3
-coating (B5.2 nm) can effectively prevent the
transition metal dissolution of the spinel-like cathode and
maintain the structure integrity of SNPCs (B0.47 nm)
(Fig. 9i). Consequently, the 1 wt% AlF
3
-modified LNMO cathode
delivered enhanced cycling stability with a high-capacity reten-
tion of around 81.7% after 100 cycles at 0.2 C and 55 1C,
whereas the pristine LNMO cathode only delivered 70.1%
capacity retention.
3.1.3. Tuning SNPCs in MOF- and COF-based cathodes.
Metal–organic frameworks (MOFs) are another type of popular
cathode materials applied in alkali metal ion rechargeable
batteries. After decades of investigation, more than twenty
thousand types of MOFs have been reported.
576
MOFs have
structures with metal ion centers (Fe, Co, Ni, Cu, Mn, etc.) and
organic ligands (polyamines, carboxylates, hydroxyl, etc.),
which give the advantages of tenable porous structures,
high surface areas, and homogeneous metal sites to these
materials and thus benefit the fast kinetic and ion diffusion
(Fig. 7c).
577,578
However, this structure also presents significant
drawbacks. First, MOFs cathode usually deliver a low specific
capacity and limited electric conductivity due to limited redox-
active sites that are only based on the metal center (M
(n+1)+
/M
n+
)
in the framework. Second, the porous framework not only
benefits the ion insertion/extraction by providing sub-
nanochannels but also leads to easy structure collapse during
the charge/discharge and low volumetric energy density. Third,
the existence of crystalline water and lattice vacancies
inside the MOFs can lead to low coulombic efficiency and poor
cycling stability.
The first MOF-based cathode material for LIBs was MIL-
53(Fe) reported by Ferey et al. in 2007. The MIL-53(Fe)
cathode provided a low capacity of around 70 mA h g
1
by
only exploiting the Fe
3+
/Fe
2+
redox reduction inside the
frameworks.
358
After that, researchers endeavor to increase
the redox-active sites and stabilize the SNPC of MOFs by
designing the organic ligands, selecting the metal node, and
modifying the nanostructure.
The organic ligand design aims to create redox-active sites
on the organic linkers of the MOFs and maintain the structure
stability simultaneously. For example, Peng et al. used tricar-
boxytriphenyl amine (TCA) to replace the conventional
quinone-type ligands and form the Cu-TCA cathode for LIBs
(Fig. 10a).
579
In addition to the redox change of Cu
2+
/Cu
+
in the
center of MOFs, the N atoms in TCA also provide redox-active
sites by transitioning between neutral and cationic forms in the
voltage range from 3.8 to 4.3 V. In the meantime, the TCA
ligands also offered good stability to support the frameworks
and sub-nanochannels (B0.5 nm).
The metal node design involves heteroatom doping and
high entropy design. The heteroatom doping can effectively
change the electronic state of the original metal node, promote
the ion transport inside sub-nanochannels, stabilize the frame-
works, and enhance the cycling voltage. For instance, Huang
et al. successfully doped Ni ions into the Prussian blue analo-
gue (PBA) K
2
FeFe(CN)
6
to form KNi
0.05
Fe
0.95
Fe(CN)
6
with a sub-
nanochannel of around 0.5 nm, which effectively changed the
electronic state of the Fe ions and enhanced K ion diffusion
inside the channels (Fig. 10b).
580
The Ni substitution promoted
the redox reaction of Fe
2+
C
6
/Fe
3+
C
6
and thus improved the
capacity from 40 to 53 mA h g
1
at high charge voltage plateaus.
Consequently, their enhanced MOF-based cathode provided a
Fig. 7 Illustration of the material structure with SNPC and the improvement strategies for (a) layered oxide cathodes, (b) polyanionic cathodes, (c) metal
organic frameworks (MOF), and (d) covalent organic framework (COF) cathode materials.
Chem Soc Rev Review Article
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2024
stable cycling performance with 83.1% capacity retention after
300 cycles at 0.1 A g
1
. A high entropy design acts the same way
as heteroatom doping but with five or more elements sharing
the same lattice sites, which further stabilizes the ion diffusion
pathways and framework structure. Ma et al. reported a high-
entropy design on the PBA (Na
x
(FeMnNiCuCo)–[Fe(CN)
6
]) with
the sub-nanochannels of around 0.6 nm and employed it as the
cathode in SIBs (Fig. 10c).
581
The high entropy (HE) structure
provided a nearly zero-strain operation during the desertion/
insertion of Na
+
ions, and thus offered a better structure
stability. As a result, the HE–BPA cathode delivered 94%
capacity retention after 150 cycles at 0.1 A g
1
.
There have been relatively fewer studies on modifiying
thestructure morphology of MOF cathodes than the above
two strategies, but it can effectively increase the intrinsic
electronic conductivity, provide a much higher density of
redox-active sites, promote ion diffusion, and offer stable sub-
nanochannels for the intercalation of large-size alkali metal
ions. For example, Wada et al. reported a 2D-MOF cathode
material (NiDI) applied in LIBs.
582
Their 2D-MOF cathode has a
bis(diimino)nickel framework with a high density of redox-
active sites and several redox states provided by both organic
ligands and metal ions. Especially, the energy storage mecha-
nism in this 2D-MOF cathode involves the insertion/desertion
Fig. 8 (a) Illustration of Mg doping into CoO
2
layers. (b) The electron density differences of NaCoO
2
(up) and Na(Co
0.92
Mg
0.08
)O
2
(down). Reproduced
from ref. 552 with permission from The Royal Society of Chemistry, Copyright 2015. (c) HRTEM image of the Na
0.8
Li
0.27
Mn
0.68
Ti
0.05
O
2
Ti-doping cathode.
Reproduced from ref. 553 with permission from American Chemical Society, Copyright 2020. (d) Schematic of the migration of Sn during the charge and
discharge in the first cycle. Reproduced from ref. 554 with permission from American Chemical Society, Copyright 2022. (e) Illustration of the structure
change during the charge of the typical NaMO
2
(left) and high entropy NaMO
2
(right), where the different colors of balls represent different elements and
the different sizes of ball represent the different oxidation states. (f) Schematic of the change between P3 and O3 phase of high entropy NaMO
2
during
the discharge. Reproduced from ref. 561 with permission from Wiley-VCH, Copyright 2020.
Review Article Chem Soc Rev
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
This journal is © The Royal Society of Chemistry 2024 Chem. Soc. Rev.
of both cations (Li
+
) and anions (PF
6
) (Fig. 10d), providing a
high specific capacity of around 155 mA h g
1
at a current
density of 10 mA g
1
. Moreover, Jiang et al. designed a 2D-MOF
cathode with a copper–benzoquinoid framework (Cu-THQ
MOF) and sub-nanochannels (B1 nm) using a similar energy
storage mechanism with cation and anion insertion/desertion
(Fig. 10e).
583
Due to the highly porous structure and intrinsic
redox characteristics of Cu-THQ, their 2D MOF cathode pro-
vides abundant and highly conductive channels for promoting
Li
+
ion transport. As a result, the 2D Cu-THQ MOF achieved a
Fig. 9 (a) Illustration of doping K
+
ions into transport ion sites in Na
3
V
2
(PO
4
)
3
. (b) The volume shrinkage between the charge and discharge of
Na
3
V
2
(PO
4
)
3
/C with different K contents. Reproduced from ref. 567 with permission from The Royal Society of Chemistry, Copyright 2014. (c) Illustration
of doping Mn
+
ions into transition metal sites in Na
3
V
2
(PO
4
)
3
. (d) HRTEM image of the Mn doped Na
3
V
2
(PO
4
)
3
with 0.373 nm SNPC. Reproduced from ref.
568 with permission from The Royal Society of Chemistry, Copyright 2016. (e) Schematic of the synthesis process for the NVPF@CNT-1 composites.
(f) TEM image of the NVPF@CNT-1 composites. Reproduced from ref. 569 with permission from Elsevier B.V., Copyright 2023. (g) Illustration of the
Na
0.877
MnSnO
4
structure (h) Schematic showing the bond lengths of (Mn
1
/Sn
1
)O
6
and (Mn
2
/Sn
2
)O
6
octahedra inside the Na
0.877
MnSnO
4
.by viewing along
the a-axial in Figure. (g) Reproduced from ref. 572 with permission from Springer Nature, Copyright 2020. (i) TEM images of the LNMO cathode with 1.0%
AlF
3
-coating. Reproduced from ref. 573 with permission from Elsevier B.V., Copyright 2020.
Chem Soc Rev Review Article
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2024
very high capacity of 387 mA h g
1
during the second cycle and
maintained a stable cycling performance with 340 mA h g
1
capacity retention after 100 cycles at 50 mA g
1
.
Covalent organic frameworks (COFs) are similar to MOFs,
but instead of having metal nodes inside the framework, COFs
are formed by covalently bonded organic ligands (Fig. 7d).
COFs, as a kind of crystalline porous material, possess abun-
dant well-defined directional meso- and nano-size channels for
ion transport, rich redox-active sites, and a stable framework
structure, offering them excellent ion diffusion capacity.
However, the drawbacks of limited specific capacity and poor
electronic and ion conductivity similar to MOFs pose chal-
lenges to the application of COFs as cathodes in alkali metal
ion batteries. The COF material was first reported by Yaghi’s
group in 2005 and named COF-1 and COF-5.
584
After ten years,
Jiang’s group first designed the COF-based cathode applied in
Li-ion batteries by growing mesoporous COFs (D
TP
-A
NDI
-COF
with channel size of around 5.06 nm) on carbon nanotubes
(CNT).
53
However, their cathode only delivered a capacity of
67 mA h g
1
at 200 mA g
1
.
Fig. 10 (a) Illustration of the Cu-TCA structure. Reproduced from ref. 579 with permission from American Chemical Society, Copyright 2016. (b)
Schematic of the crystal structure of the K
2
FeFe(CN)
6
(up) and Ni doping KNi
0.05
Fe
0.95
Fe(CN)
6
(down). Reproduced from ref. 580 with permission from
American Chemical Society, Copyright 2019. (c) Schematic of the high entropy Na
x
(FeMnNiCuCo)–[Fe(CN)
6
] crystal structure. Reproduced from ref. 581
with permission from Wiley-VCH, Copyright 2021. (d) Illustration of the energy storage mechanism in the NiDI cathode. Reproduced from ref. 582 with
permission from Wiley-VCH, Copyright 2018. (e) Schematic of the 2D Cu-THQ-MOF structure. Reproduced from ref. 583 with permission from Wiley-
VCH, Copyright 2019.
Review Article Chem Soc Rev
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
This journal is © The Royal Society of Chemistry 2024 Chem. Soc. Rev.
In recent years, several strategies, such as SNPC design, pore
wall decoration, and morphology modification (exfoliation,
crystallinity regulation and orientation, secondary material
combination, etc.) have been developed to improve the perfor-
mance of COF-based cathodes. Generally, the large pores can
provide more free space for ion transport, and small pores can
be more conductive due to the interaction between the pore
wall and transport ions. Combined with the development of
enriching redox-active centers on the pore walls, COF-based
cathodes can offer significantly enhanced electric and ionic
conductivities and excellent electrochemical performance. For
example, Shi et al. developed a nitrogen-rich COF with triqui-
noxalinylene and benzoquinone units (TQBQ-COF).
355
The
TQBQ-COF has a honeycomb-like and AB stacking structure
with well-defined directional SNPCs of around 1 nm (Fig. 11a
and b). The SNPC provided the TQBQ-COF with a high electro-
nic conductivity of 1.973 10
9
Scm
1
, and the abundant N
atoms on the pore walls effectively decrease the energy gap
between the lowest and highest occupied molecular orbitals,
thus promoting the ionic and electronic conductivity. Conse-
quently, the TQBQ-COF cathode in SIBs provided a very high
capacity of 327.2 mA h g
1
and around 89% capacity retention
after 400 cycles at 0.1 A g
1
.
The most popular structural morphology modification of
COFs is the 2D morphology obtained through direct synthesis
or exfoliating from 3D COFs, which can effectively enlarge their
surface area, improve the accessibility of redox-active sites, and
promote ion diffusion. For example, Wang et al. exfoliated the
anthraquinone-based COF (DAAQ-TFP-COF) to redox-active
ECOF (DAAQ-ECOF), providing an approximately three times
faster Li
+
diffusion rate than its non-exfoliated counterpart due
to the significantly shortened ion transport paths (Fig. 11c).
585
Wu et al. reported a 2D COF cathode (BQ1-COF) through the
polycondensation reaction between C
6
N
2
O
2
and C
8
N
4
O
2
to
deliver CRO and CRN groups (Fig. 11d).
353
This design
provided SNPCs of B1 nm and minimized the inactive group
inside the BQ1-COF (Fig. 11e). As a result, the BQ1-COF cathode
in LIBs offered an excellent reversible capacity of around
502 mA h g
1
at 0.05C and very stable cycling performance
with 81% capacity retention after 1000 cycles at 1.54 A g
1
.
It is worth mentioning that conjugated microporous poly-
mers (CMPs), from some perspectives, can be regarded as
amorphous COFs. They are another type of microporous mate-
rial with promising potential as cathode materials in alkali
metal ion batteries.
588
CMPs have a very similar composition to
COFs, but instead of a crystalline structure, CMPs are amor-
phous. This characteristic not only makes the CMPs more
challenging to characterize and process than COFs but also
gives CMPs the advantages of high thermal and chemical
stability. Moreover, CMPs have a highly conjugated structure
and extended p-conjugation system, which benefits the stability
of the framework and the ion transport in SNPCs. For instance,
Li et al. designed a CMP-based cathode applied in SIBs by cross-
linking 1,8-diaminopentacene-5,7,12,14-teraone (DAPT) and
1,3,5triformylphloroglucinol (TFP) to form p-conjugated porous
frameworks (DAPT-TFP-CPFs) (Fig. 11f).
586
The DAPT-TFP-CPF
cathode resulted in fast ionic diffusion thanks to the sub
nanometer pores (B0.8 nm) and the lamellar structure.
Furthermore, the extended p-conjugation backbones effectively
stabilized the SNPC inside the framework, improving the
charge transport stability in cycling. As a result, the DAPT-
TFP-CPF cathode exhibited a very stable cycling performance
with a high-capacity retention of 145 mA h g
1
after 1000 cycles
at 100 mA g
1
.
It is worth noting that, in addition to monovalent (Li
+
,Na
+
,
and K
+
) ion batteries, tuning of SNPCs in cathodes of multi-
valent ion (Zn
2+
,Mg
2+
,Ca
2+
,Al
3+
) batteries can also remarkably
enhance their electrochemical performance. Compared with
monovalent ion batteries, multivalent ion batteries offer advan-
tages including more affordable cost, higher safety and lower
environmental impact. However, challenges such as unstable
electrode materials, poor ion diffusion kinetics, and a narrow
electrochemical stable potential window hinder the application
of multivalent ion batteries. Tuning SNPCs in the cathode
materials, through methods such as the abovementioned het-
eroatom doping, size regulation, pre-intercalation, defect engi-
neering, and freestanding configuration design, could pave the
way for breakthroughs and potential applications in these
multivalent ion batteries.
589
For instance, Zhu et al. designed
a freestanding design Mn-based cathode (SSWM@Mn
3
O
4
) for
Zn-ion batteries using stainless steel welded mesh (SSWM) to
grow the flower-like Mn
3
O
4
(Fig. 11g and h). The cathode
provided SNPCs of around 0.61 nm (Fig. 11i), allowing Zn
2+
intercalation/extraction and effectively improving the cycling
stability up to 500 cycles at 500 mA g
1
.
587
Nam et al. intro-
duced crystal water into the SNPCs (B0.7 nm) inside the
layered manganese oxide to form an enhanced stable cathode
for Mg-ion batteries at high voltage (2.8 V vs. Mg/Mg
2+
).
590
The
Mg-ion batteries with their cathode offered 62.5% capacity
retention after 10 000 cycles. Furthermore, Li et al. reported a
Co
3
S
4
microsphere cathode material for Al-ion batteries.
591
The
Co
3
S
4
microsphere provided about 0.2 nm SNPCs for Al
3+
intercalation/deintercalation and a porous structure for electro-
lyte penetration. With this unique structure design, the Al-ion
batteries with Co
3
S
4
microsphere cathodes delivered a reversi-
ble capacity of 90 mA h g
1
after 150 cycles at 50 mA g
1
.
3.1.4. The design of SNPCs in conversion-type cathode
materials. Cathode materials based on conversion reactions,
such as S and selenium (Se) have been intensively investigated
in the past decades due to their high theoretical specific
capacities (Fig. 12a and b).
592,593
At room temperature, the
reactions of S/Se cathodes during charge and discharge
processes involve multiple phase-transition steps with the
formation of soluble polysulfides (PSs)/polyselenides (PSes) as
intermediates. The major challenges faced by these conversion-
type cathode materials include poor electronic conductivity of
the cathode materials, the ‘‘shuttle effect’’ of PS/PSe intermedi-
ates, significant volume variations during cation insertion and
extraction, and sluggish reaction kinetics. To solve the above-
mentioned issues, a strategy which involves introducing multi-
functional hosts with appropriate SNPC design to form S/Se-
based nanocomposites has made significant achievements so
Chem Soc Rev Review Article
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2024
far. The host materials with SNPCs, such as nanostructured
porous carbon materials, provide sub-nanoscale conductive
networks that significantly increase the electronic conductivity
of the S/Se cathodes and buffer the volume change during
cation insertion and extraction in different battery systems (e.g.,
Li/Na/K–S batteries and Li/Na/K–Se batteries). More impor-
tantly, the confinement effects of SNPC suppress the ‘‘shuttle
effect’’ (i.e., restricting the dissolution of PS/PSe intermediates
into the electrolytes), reducing the loss of active materials and
the side reactions on the anode counterparts.
Carbon-based materials with SNPCs, ranging from 1D CNTs
to 2D graphene and 3D porous carbon spheres, are the most
intensively investigated hosts for sulfur cathodes.
598–604
Com-
pared with the traditional porous structure design, carbon
materials with SNPCs have been found to show unusual elec-
trochemical reaction behaviours in the S cathode. For instance,
Fig. 11 (a) Schematic of the TQBQ-COF chemical structure, electrochemical redox mechanism, and the theoretical capacity of 515 mA h g
1
.(b)
Illustration of the AB stacking model of TQBQ-COF layers with a packing distance of 3.07 Å. Reproduced from ref. 355 with permission from Springer
Nature, Copyright 2020. (c) Schematic of the DAAQ-TFP-COF exfoliation to DAAQ-ECOF as cathodes for LIBs. Reproduced from ref. 585 with
permission from American Chemical Society, Copyright 2017. (d) Illustration of the one-layer-conjugated structure of the BQ1-COF and the elements
inside the skeleton. (e) TEM images of the BQ1-COF showing the interlayered spacing between the adjacent 2D COF nanosheets. Reproduced from ref.
353 with permission from Elsevier Ltd, Copyright 2020. (f) Schematic of the synthesis route and structure of DAPT-TFP-CPF. Reproduced from ref. 586
with permission from American Chemical Society, Copyright 2018. (g) Schematic synthesis process of the SSWM@Mn
3
O
4
cathode material. (h) SEM
image of SSWM@Mn
3
O
4
. (i) HRTEM image of the SSWM@Mn
3
O
4
cathode showing the SNPCs. Reproduced from ref. 587 with permission from The Royal
Society of Chemistry, Copyright 2018.
Review Article Chem Soc Rev
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
This journal is © The Royal Society of Chemistry 2024 Chem. Soc. Rev.
two types of single-walled carbon nanotubes (SWCNTs) with
different diameters were prepared as hosts to study the 1D
confinement effects of SNPCs on the lithiation/delithiation
process of S cathodes in Li–S batteries (Fig. 12c).
594
The long-
chain S diradicals were formed inside the as-prepared SWCNTs
via infusion and polymerization processes. The SWCNTs pre-
pared by the electric arc method (EA-SWCNTs) have an average
diameter of B1.55 nm with an average inner van der Waals
diameter of 12.1 Å for EA-SWNTs. The other SWCNTs obtained
from high-pressure carbon monoxide (HiPc0-SWCNTs) have an
average diameter of B1.0 nm with an average inner van der
Waals diameter of 6.6 Å for HiPco-SWNTs. The conventional
solid–liquid–solid reactions occurred in the S cathode with
EA-SWCNTs in both tetraethylene glycol dimethyl ether
(TEGDME)-based and 1,4,7,10,13-pentaoxacyclopentadecane
(15-crown-5)-based electrolytes. In contrast, unusual continu-
ous solid-state reactions were observed in S cathodes with
HiPc0-SWCNTs. The distinct electrochemical behaviors were
explained because of the unique SNPC feature of HiPc0-
SWCNTs. The largest van der Waals dimensions of the solvated
[Li(TEGDME)]
+
and [Li(15-crown-5)]
+
ions are 10.87 Å and
11.34 Å, respectively, which are smaller than the diameter of
EA-SWCNTs but larger than that of HiPc0-SWCNTs. Therefore,
the S cathode with EA-SWCNTs showed similar phase transfor-
mation processes to the normal S cathodes as the solvated
[Li(TEGDME)]
+
and [Li(15-crown-5)]
+
ions could still access the
inner channels of EA-SWCNTs and react with S. However, the
sulfur inside HiPc0-SWCNTs could not directedly react with
the solvated cations. Instead, the sulfur was reduced through
the wall of SWCNTs via an out-of-plane p-electron interaction.
A 2D graphene material with an SNPC structure (i.e.,an
interlayer spacing of B0.4 nm) was developed as the S host for
Li–S batteries. Through a solvothermal method or an interla-
mellar reaction method, S molecules were intercalated into the
graphene interlayers to form a S/reduced graphene oxide
(S/RGO) composite (Fig. 12d). The electrochemical performance
of the S/RGO composite showed similar phase-transfer beha-
viors to that of small S molecules in a carbonate-based electro-
lyte (1 M LiTFSI in DOL/DME), demonstrating that this 2D
SNPC design effectively facilitates phase transformation
during electrochemical conversion reactions from S to lithium
sulfide.
595
Another carbon host with 3D SNPC design, micro-
mesoporous carbon nanospheres (MMPCS), was reported by
Wu et al. for RT Na–S batteries.
596
Porous carbon spheres with
tailored continuous carbonaceous pores ranging from micro-
meters to sub-nanometers were synthesized as S hosts
(Fig. 12e). Similarly, the design of the 3D SNPC carbon host
has been widely applied in the Li/Na–Se battery system. Aboo-
nasr Shiraz et al. synthesized microporous carbon (MPC) with
SNPCs of around 1 nm from the PVDF precursor. When
employed as a host for Se in Li–Se batteries, the MPC could
confine the LiPSes and effectively reduce the side reaction
between the Se cathode and the electrolyte (Fig. 12f).
366
As a
result, the Li–Se batteries exhibited a reversible capacity of
508.8 mA h g
1
after 100 cycles at 0.1C. Furthermore, Zhang
et al. designed interconnected conductive polyaniline (PANI)
coated hierarchical porous carbon (i-PANI@NSHPC) to accom-
odate up to 65.7 wt% Se in SNPCs applied as cathodes in Na–Se
batteries.
595
This cathode offered an excellent reversible capa-
city of 617 mA h g
1 at 0.2C after 200 cycles due to the
enhanced electronic and ionic conductivity even at high Se
areal mass loading. The S material confined within those
multifunctional channels is highly reactive with full access to
Na
+
ions. In addition, a highly Na-ion conductive and contin-
uous solid interphase containing sodium sulfide was confor-
mally formed inside the pores, significantly stabilizing the
cathode/electrolyte interfaces, and promoting the kinetics of
S-involved redox reactions. In addition to the carbon-based
materials, MOF and COF 3D frameworks with SNPCs have also
been used as hosts for S/Se cathodes. For instance, Su et al.
reported a 3D SNPC framework by using the Prussian blue
analogues (PBAs) coated with PEDOT (Na
2
Fe[Fe(CN)
6
]@
poly(3,4-ethylenedioxythiophene)) to capture the sulfur and
polysulfides for Li–S batteries (Fig. 12g).
597
The MOF-based
host significantly enhanced the confinement of soluble Li
polysulfides through the Lewis acid–base interaction between
the rich open metal sites in the SNPC of the MOF host and
polysulfide anions. Finally, their S@Na
2
Fe[Fe(CN)
6
]@PEDOT
cathode exhibited a stable cycling performance with a capacity
retention of 697 mA h g
1
after 100 cycles at 2C. Recently,
redox-active polymers have emerged as an attractive option
among organic materials for seving as S hosts, offering advan-
tages such as cost-effectiveness, diversity, good processability,
unique electrochemical properties, and precise tuning for
energy storage applications.
605
e.g., Park et al. the reported S-
rich polymers on the gram scale derived from the vulcanization
of trithiocyanuric acid-based materials. The resulting cathode
material exhibited stable cycling over 450 cycles, with a reten-
tion rate of 83% and excellent rate performance with a current
density of C/10 (1210 mA h g
1
) to 5C (730 mA h g
1
). This
performance is attributed to strong bonding energy between
redox-reactive polymers and S.
606
In conclusion, the SNPCs play an important role in both
intercalation-type and conversion-type cathode materials. The
size, distribution, and amount of SNPCs inside cathode materi-
als directly affect the accommodation and transportation of
metal ions and the confinement of active materials. The
advanced material designs to modify the morphology and
catalytic functions of the SNPCs in cathodes can further
improve the electrochemical performance of rechargeable
batteries.
3.2. Electrochemical couplings of anode materials with
SNPCs
In addition to their significant role in cathode materials, SNPCs
also play critical roles in various anodes. It has been widely
reported that the SNPCs can be engineered on the surface of the
anodes as artificial solid–electrolyte interphase (SEI) layers,
which could facilitate feasible selective ion transport, avoid
excessive SEI formation, and alleviate the volume changes upon
cycling. The SNPCs can also be engineered inside the nano-
particles of the anode materials, which could generate fast ion
Chem Soc Rev Review Article
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2024
transportation to achieve enhanced rate performance. Hereby,
the rational design of SNPCs for various types of anode materi-
als has been systematically sorted and discussed.
3.2.1. Intercalation-based anodes. The intercalation pro-
cess to insert diversified working ions into the interlayers of the
crystal planes represents a wide variety of anode materials for
versatile battery systems, of which the intercalation of Li
+
into
the graphite layers has been the fundamental working mecha-
nism for commercial LIBs since 1990s. Carbonaceous materi-
als, conjugated microporous polymers CMPs, titanates, TMDs,
and Mxenes generally working as intercalation-based anodes,
which exhibit vital importance in practical batteries due to their
Fig. 12 (a) Schematic illustration showing the mass/charge transfer at the cathode–electrolyte interface in liquid-electrolyte-based Li–S batteries. S is
reduced to generate LiPSs, which could dissolve in the liquid electrolyte and shuttle between two electrodes. Reproduced from ref. 592 with permission
from Wiley-VCH, Copyright 2017. (b) Typical charge/discharge profile of Li–S batteries and chemical evolution of LiPS species at different states of
(dis)charge. Reproduced from ref. 593 with permission from Wiley-VCH, Copyright 2015. (c) S@SWNT viewed along the axis and side length as computed
with dispersion-corrected DFT. Reproduced from ref. 594 with permission from American Chemical Society, Copyright 2018. (d) Schematic illustration of
the fabrication processes of the sulfur intercalated reduced graphite oxide by two wet chemical methods. Reproduced from ref. 595 with permission
from Wiley-VCH, Copyright 2016. (e) Schematic illustration of Na
+
transmission in the internal structures of MMPCS-800@S. Reproduced from ref. 596
with permission from Wiley-VCH, Copyright 2022. (f) Microporous carbon (MPC) with SNPCs as a host for the Se cathode. Reproduced from ref. 366 with
permission from Elsevier B.V, Copyright 2019. (g) Atomic model configurations showing the interactions between Na
2
Fe[Fe(CN)
6
] and polysulfide Li
2
S
x
(x= 8, 6, 4, and 2). Reproduced from ref. 597 with permission from Wiley-VCH, Copyright 2017.
Review Article Chem Soc Rev
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
This journal is © The Royal Society of Chemistry 2024 Chem. Soc. Rev.
outstanding cycling performance. These anodes with SNPCs
could deliver enhanced electrochemical performance, as a
result of the benefits derived from the SNPCs.
3.2.1.1. Carbon-based anodes. Carbon-based materials
including graphite, graphene, graphdiyne, hard carbon, soft
carbon, carbon nanotubes (CNTs), etc. are all promising anode
materials for alkali metal ion batteries. Numerous research
studies have incorporated SNPCs into these carbonaceous
anodes to achieve outstanding electrochemical performance.
For instance, hard carbon has been widely recognised as a
versatile anode for alkali metal batteries, the abundant SNPCs
in the hard carbon resulting from randomly stacked graphitic
layers render it capable of accommodating more ions both
between the layers and also in the sub-nano pores, as presented
in Fig. 13.
607
It is worth noting that, the SNPCs in the hard
carbon anodes could contribute a considerable amount of low
voltage capacity, which is achieved probably by pore-filling of
quasi-metallic Na clusters.
608
The size of the SNPCs in the
carbon anodes is crucial for the accommodation of alkali metal
ions, especially for K
+
with a larger diameter. For instance, a
nitrogen-doped porous carbon (NPC) has been reported for
KIBs, the SNPC size was enlarged from 0.335 to 0.383 nm due to
the decomposition of the Na
2
CO
3
during pyrolysis, as presented
in Fig. 13b.
609
As a result, the NPC delivered a high reversible
capacity of 382 mA h g
1
for more than 500 cycles and good rate
capability of 185 mA h g
1
at 10.0 A g
1
. The improvements in
the performance were attributed to the surface driven K
+
storage by the enlarged SNPCs, as well as enhanced K
+
affinity
provided by N-doping sites. Similarly, it has been reported that
the interlayer spacing of the graphitized domains in the hard
carbon can be expanded to 0.42 nm by incorporating graphene
oxide in the precursors; the expanded spacing could serve as
fast K
+
diffusion channels.
394
As a result, high reversible
capacity, excellent rate capability, and good cycling stability
have been achieved for PIBs.
Alternatively, Mai and colleagues reported soft carbon
nanosheets with SNPCs of 0.45–0.90 nm, exhibiting a high
surface area, large pore volume, and rich defective sites.
610
Excellent capacitance-dominated Na
+
and K
+
storage was
achieved, as a result of the extra ion storage sites derived from
abundant SNPCs. Significantly, Li and co-workers reported the
synthesis of graphdiyne with a channel of 0.546 nm as anode
materials for LIBs, which could be expanded to 0.596 nm upon
Li
+
intercalation, as illustrated in Fig. 13c.
611
The self-
expanding channel could reduce the energy barrier and elec-
trostatic repulsion for Li
+
transport, facilitating fast solid-state
Li
+
diffusion. Benefiting from self-expanded SNPCs, graphdiyne
delivered higher capacity, better cycling stability, enhanced
capacity retention, and broaden working temperature range
at high current rates. Besides, S-doped hollow carbon
nanosheets,
612
hard carbon,
613
and N,P co-doped hollow micro-
porous carbon,
614
have also been reported to deliver
enhanced electrochemical performance due to the incorpora-
tion of SNPCs, as a result of enhanced ion transportation,
improved electrolyte wettability, and decreased SEI thickness.
Furthermore, the advantage of the SNPCs in various carbo-
naceous materials including graphite nanomesh,
615
holey
graphite,
616
and microporous carbon
617
have also been
revealed via theoretical calculations, further demonstrating
the significance of SNPCs for various carbon anodes.
3.2.1.2. Titanates, TMDs, and MXenes. Titanates are widely
used intercalation-based anode materials for alkali metal ion
batteries, which exhibit good cycling stability and excellent rate
performance.
618
Various TMDs and MXenes with a well aligned
layer structure could also deliver excellent intercalation perfor-
mance for rechargeable batteries, benefiting from the stable
layered structure with SNPCs. Previous research by Geng and
co-workers reported stacked titania sheets with pillared inter-
layer spacing ranging from 0.76 to 1.15 nm, which could
facilitate ultrafast intercalation of Li
+
,Na
+
and K
+
with high
cycling stability, as presented in Fig. 13d.
619
The ultrafast ion
transportation could be attributed to the open 2D SNPCs,
which could even achieve good rate performance for PIBs with
a thick electrode of up to 80 mm. Besides, several research
groups reported MoS
2
and MoO
3
based TMDs for LIBs and
sodium-ion batteries (SIBs) with SNPCs of B1 nm, which
delivered excellent cycling stability and rate capability.
620,621
The improved electrochemical performance can be attributed
to SNPCs between the layered structures, which could acceler-
ate ion transportation and also decrease the ion diffusion
pathway. A wide variety of Mxenes with layered structures
and abundant SNPCs have also been widely used as anode
materials. For instance, Yamada and colleagues reported a TiC
2
MXene for Na
+
storage in hybrid capacitors and SIBs, which
could retain a stable SNPC of 1.01 nm after the first activation
process as a result of the Na
+
pillars, as illustrated in Fig. 13e.
388
The SIB full cells delivered a high specific energy of
260 W h kg
1
at a high specific power of 1.4 kW kg
1
, as a result
of the rich and stable SNPC structures. Similarly, a pillared
MXene with an expanded interlayer pacing of 1.02 nm has been
reported for LIBs, which could enhance the specific capacity,
rate capability, and also the cycling stability.
622
The improved
performance was attributed to SNPCs, which could accelerate
ion transportation and provide more ion storage sites. The
benefits of the SNPCs for various intercalation based anodes
were further verified by numerous previous reports, demon-
strating the universality of the SNPCs for layered anodes in
delivering enhanced rate performance as a result of the boasted
ion transportation in the confined channels between the
layers.
623–625
3.2.2. COFs. The structure of COFs could enable good
electron transport in conductive polymer backbones, as well
as effective ion transport through SNPCs. Therefore, COFs with
redox active building blocks to accommodate versatile metal
ions in SNPCs have been frequently reported as anode materi-
als for various batteries, working with a similar mechanism to
that of the Li
+
intercalation into the graphite anode. Among
which, CMP is a distinctive family of COFs, featuring extended
p-conjugation backbones and permanent nanopores.
626,627
As
presented in Fig. 13f, Chen and colleagues reported a
Chem Soc Rev Review Article
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2024
phenazine based CMP with rich SNPCs, which could host
various mono- and multi-valent charge carriers (H
+
,Li
+
,Na
+
,
K
+
,Zn
2+
, and Al
3+
) for diverse aqueous rechargeable batteries
combining rapid kinetics, ultralong lifespan, and chemical
rechargeability.
543,628
The highly reversible redox activity, facile
intramolecular electron transfer, and high ion diffusion coeffi-
cient were attributed to the high density of accessible redox
sites derived from the cross-linked structure with rich SNPCs.
Alternatively, a 2D nitrogen rich CMP with regular SNPCs of 0.7
and 1.1 nm has been rationally designed for LIBs, delivering a
high reversible capacity of 701 mA h g
1
at 1 A g
1
and
remarkable cycling stability for more than 500 cycles.
629
The
superior rate performance can be attributed to excellent con-
ductive backbones, as well as the reduced Li
+
diffusion pathway
derived from the SNPCs. Several excellent reviews have system-
atically summarized the synthesis, properties, and applications
of various COFs,
34,44,49,116,626,627
which should be enlightening
for follow up research on rational design of high performance
CMP anode materials for various rechargeable batteries.
3.2.3. Conversion-based anodes. Conversion based anodes
are promising candidates for rechargeable batteries with high
specific capacity, whereas suffering from poor cycling stability
and rapid capacity degradation.
630,631
Recent reports demon-
strate that the drawbacks of the conversion based anodes can
be significantly relieved by incorporating with SNPCs. As pre-
sented in Fig. 14a, ultrathin Co
3
O
4
nanosheets with an inter-
layer spacing of 0.6 nm have been reported for LIBs, which
delivered high specific capacity (1230 mA h g
1
at 0.2 A g
1
),
Fig. 13 (a) Schematic illustration of the accommodation of Li
+
/Na
+
/K
+
in hard carbon anodes through intercalation, adsorption, and pore filling in the
SNPCs.
603
Reproduced from ref. 603 with permission from Wiley-VCH, Copyright 2019. (b) Schematic illustration of the synthesis of nitrogen doped
porous carbon and its working mechanism for K
+
storage.
594
Reproduced from ref. 594 with permission from Wiley-VCH, Copyright 2018. (c) Illustration
of the working mechanism of graphdiyne for Li
+
storage.
596
Reproduced from ref. 596 with permission from Wiley-VCH, Copyright 2018. (d) Schematic
illustration of the intercalation of Li
+
/Na
+
/K
+
in titania sheets.
610
Reproduced from ref. 610 with permission from Wiley-VCH, Copyright 2019. (e)
Schematic illustration of the synthesis of MXene and the structural stability of the SNPC upon Na
+
intercalation/deintercalation.
387
Reproduced from ref.
387 with permission from Nature Publishing Group, Copyright 2015. (f) Schematic illustration of the synthesis of CMP as a universal and ultra robust
organic electrode for diverse aqueous battery chemistry.
611
Reproduced from ref. 611 with permission from Wiley-VCH, Copyright 2018.
Review Article Chem Soc Rev
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
This journal is © The Royal Society of Chemistry 2024 Chem. Soc. Rev.
excellent rate performance, and excellent cycle capability (1500
cycles).
632
The enhanced performance can be attributed to the
SNPCs between the layers, which could establish 2D nanoflui-
dic channels offering extra lithium storage sites, accelerate
horizontal and vertical Li-ion transport, and alleviate volume
changes during the cycling. Alternatively, Park and colleagues
reported an exfoliated 2D TMD (MoS
2
) nanosheets with inter-
layer channels of E1 nm for LIBs, which delivered a high
specific capacity of 933.1 mA h g
1
at 0.1 A g
1
and excellent
cycling stability with 90% retention after 1000 cycles.
633
The
SNPCs between the layers with good tolerance to volume
expansion and negatively charged channels for fast ion trans-
portation were attributed to the enhancement of the electro-
chemical performance. It has also been reported that Ni
2
P
anchored in nitrogen doped carbon (Ni
2
P@N–C) with rich
SNPCs could deliver enhanced rate capability (197 mA h g
1
at 2 A g
1
) and excellent cycling stability (0.06% capacity decay
per cycle), as a result of the high electrical conductivity of the
matrix and confinement effect of the SNPCs (Fig. 14b).
634
The
advantages of SNPCs have also been reported for other
conversion-based anodes including silicon oxide,
635
SnO
2
,
636
and others high capacity conversion-based anodes with SNPCs
can be expected in the following research.
3.2.4. Alloy-based anodes. Alloy based anodes including Si,
Ge, Sn, P, Sb, Bi, etc. are well known for their ultrahigh specific
capacities resulting from alloying reactions, while their prac-
tical applications have been seriously impeded as a result of the
huge volume change, unstable SEI layers, and poor cycling
stability.
637–639
Confining the alloy based anodes in SNPCs has
been reported to be an effective approach to address these
issues. For instances, Lou and colleagues reported a COF based
artificial SEI layer coating on Si nanoparticles, the anode
delivered a high specific capacity of 1864 mA h g
1
at a high
current density of 2000 mA g
1
and a high capacity retention of
more than 60% after 1000 cycles.
640
The COF based artificial
SEI with SNPCs could facilitate fast ion transportation, reduced
electrolyte decomposition, enhanced Coulombic efficiency, as
well as improved cycling stability, as illustrated in Fig. 14c.
Alternatively, carbon coating layers on Si with abundant SNPCs
derived from pyrolyzed metal–organic frameworks (MOF) have
been reported for LIBs, resulting in high specific capacity and
exceptional cycling stability.
403
The increased electrical con-
ductivity, relieved volume change, and enhanced ionic trans-
portation derived from the SNPCs are all responsible for the
dramatically enhanced electrochemical performance of the Si
anode. Similarly, Yu and coworkers reported a MOF-derived
SNPCs (o1 nm) rich carbon to accommodate amorphous Red P
as anode material for SIBs (Fig. 14d), the composite anode
delivered a high specific capacity of 600 mA h g
1
and excellent
cycling performance for 1000 cycles.
405
The enhanced perfor-
mance can be attributed to the buffering effect, enhanced ionic
conductivity, and improved electronic conductivity derived
from the SNPCs. Various recent research also reported the
encapsulation of red P into SNPCs to achieve dramatically
enhanced electrochemical performances.
404,406–408
With attrac-
tive size, buffering, and confinement effects, the enormous
potential of SNPCs for various other kinds of conversion-based
anodes should be further explored.
3.2.5. Metal anodes. The ultimate choice of anode materi-
als for the rechargeable batteries should be metal anodes, i.e.,
Li, Na, K, and Zn metal anodes for LIBs, SIBs, PIBs, and
aqueous zinc ion batteries (AZBs), respectively.
641
For instance,
Li metal anodes could deliver the highest theoretical capacity of
3860 mA h g
1
for LIBs as compared to all other kinds of
anodes.
642,643
Whereas metal anodes exhibit all the problems of
other types of anodes, and the extent is much more serious.
Besides, the metal anodes are prone to dendrite growth caused
by internal short circuits, posing serious safety concerns for the
practical application of the metal batteries.
3.2.5.1. Li/Na/K metal anodes. cLi metal anodes have been
considered as the ‘‘Holy Grail’’ for LIBs due to their highest
specific capacity and lowest redox potential, to Na and K metals
for SIBs and PIBs. The alkali metal anodes generally exhibit
huge volume change, unstable SEI layers, low Coulombic
efficiency, and continuous dendrite growth during repeated
cycling.
644,645
Various species with SNPCs have been reported
as artificial SEI layers to protect the metal anodes. For example,
an ACOF coating layer with rich SNPC structures has been used
for Li metal anodes (Fig. 15a), exhibiting a size effect that
enables selective Li
+
transport through the SNPCs while pre-
venting the penetration of electrolyte solvents.
646
As a result of
effective protection from the ACOF coating layers, significantly
enhanced cycling stability for Li metal anodes and Li metal full
cells has been achieved. Here, 3D porous frameworks have
attracted tremendous interest as metal anode protective layers
for rechargeable batteries because of their large active surface
areas. Yun et al. demonstrated 3D microporous frameworks
grown on Li anodes. Benefiting from the 3D microporous
framework induced interfacial activity gradient, Li ions under-
goe a uniform deposition, boosting the cycle performance of Li
metal batteries.
647
As the most representative type of material
in the 3D framework, MOFs, COFs and their derived materials
have also received a lot of attention. For instance, similarly, a
COF film with size effects to block TFSI
while allowing for the
selective permeation of Li
+
was designed as an artificial SEI
layer for Li metal anodes (Fig. 15b), which could generate
uniform Li metal plating as compared to the bare Li metal
anodes with serious dendrite growth.
648
Significantly, Man-
thiram and co-workers reported the room temperature in situ
growth of a COF layer on the Li metal anode, as illustrated in
Fig. 15c, which exhibits a suppression effect to prevent dendrite
growth, as well as a size effect to achieve uniform Li
+
flux.
649
Benefiting from the COF protective layer with rich SNPC
structures, the Li–S cells delivered exceptional cycling stability
for more than 600 cycles with a low-capacity decay of 0.05% per
cycle under lean electrolyte conditions. Besides, CMP-based
protective layers with SNPCs of 0.5–0.6 nm as ion-selective
nanofluidic transport have been reported for Li metal anodes,
which delivered extraordinary cycling stability for 2550 h at a
high areal current density of 20 mA cm
2
.
139
The dramatically
enhanced cycling performance of the Li metal anode can be
Chem Soc Rev Review Article
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2024
attributed to SNPCs with Li
+
selective channels, so as to prevent
the side reactions of Li metal with anions and electrolytes.
Similarly, MOF-based protective layers have been coated on Cu
current collectors as artificial SEI layers for Li and Na metal
anodes with excellent cycling stability.
650,651
Furthermore,
expanded graphite with SNPCs of around 0.7 nm as molecular
tunnelling has been designed as a host material for Li metal,
which could realize dendrite-free Li metal anodes and Li metal
full cells with no noticeable capacity degradation after 370
cycles.
652
After that, other MOF- and COF-derived materials
have also been successfully prepared as protective layer for
metal anodes.
653–656
The performance enhancement was attrib-
uted to the bulk diffusion of superdense Li into SNPCs. It can
be anticipated that various SNPC-based artificial SEI layers and
host materials could further enhance the cycling stability of
alkali metal anodes, paving the way for the practical application
of metal anodes for next-generation high energy density
rechargeable batteries.
3.2.5.2. Zn metal anodes. Aqueous zinc ion batteries (AZBs)
have been intensively investigated as alternatives for alkali
metal batteries with organic electrolytes, which could deliver
high specific capacity and non-flammable properties.
659
Low
Coulombic efficiency due to hydrogen evolution and short life
span as a result of Zn dendrite growth are the prevailing
challenges for AZBs. To address these issues, a zeolite based
protective layer with SNPCs of 0.349 nm has been designed for
Zn metal anodes, which could facilitate selective Zn
2+
transpor-
tation through channel size restriction and electric field repul-
sion to the anions, as presented in Fig. 15d.
657
As such, the side
reactions can be restricted to achieve enhanced Coulombic
efficiency. Besides, dendrite-free Zn metal deposition can be
achieved due to the homogenized Zn
2+
flux derived from uni-
form SNPCs. Consequently, a long lifespan of 2400 h, a high
current tolerance of 100 mA cm
2
, and a high capacity reten-
tion of 76.4% after 7500 cycles have been achieved. Similarly,
an ion regulating interface has been developed for Zn metal
anodes, which delivered a high Coulombic efficiency of 99.9%
and stable cycling for 2500 h, as a result of the homogeneous
Zn
2+
flux derived from the coating layer with rich SNPCs
(Fig. 15e).
658
Interestingly, a vermiculite (VRM) coating with
rich negatively charged SNPCs has been designed for Zn metal
anodes (Fig. 15f), which could accelerate the transport of Zn
2+
via electrostatic effects while reducing the water molecules in
the Zn
2+
solvation shell.
635
Therefore, the VRM@Zn anode
delivered excellent cycling stability for 5000 cycles at 1A g
1
.
The feasibility of protective coating layers for Zn metal anodes
could be further extended to other SNPC based materials.
Versatile anode materials incorporating SNPCs to achieve
enhanced electrochemical performance for various battery sys-
tems have been summarized in Table 5. Further exploration of
SNPCs for various anode materials is essentially required,
which holds promise for enhancing the performance of anodes
for next-generation batteries with high energy density, high
power density, and desirable cycling stability.
3.3. Electrolyte
An electrolyte is the medium which carries the ionic charges
between the electrodes. Although its role seems trivial, the
electrolyte properties drastically influence the electrochemical
Fig. 14 (a) Schematic illustration of ultrathin functionalized Co
3
O
4
nanosheets with massive arrays of 2D nanofluidic channels for Li-ion transport.
Reproduced from ref. 632 with permission from Wiley-VCH, Copyright 2017. (b) Schematic illustration of the sodiation/desodiation process for the
Ni
2
P@C-N anode. Reproduced from ref. 634 with permission from Elsevier, Copyright 2019. (c) Schematic illustration of the Si@COF anode with SNPCs
for LIBs. Reproduced from ref. 403 with permission from Elsevier, Copyright 2020. (d) Schematic illustration of the sodiation process of the red P
confined in SNPCs. Reproduced from ref. 405 with permission from Wiley-VCH, Copyright 2017.
Review Article Chem Soc Rev
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
This journal is © The Royal Society of Chemistry 2024 Chem. Soc. Rev.
battery performances. Ideally, the electrolyte should (1) be
electrochemically stable at the voltage required for oxidation
and reduction reactions at the electrodes; (2) have a high ionic
conductivity of 410
3
Scm
1
, and a cation transference
number close to 1; (3) present high safety; (4) operate over a
wide temperature range; and (5) represent a small volume or
weight fraction of the battery. Currently, in LIBs, the common
electrolytes are based on liquid organic solvent and thermally
unstable Li salts. As largely demonstrated in the literature,
these types of electrolytes are not satisfactory to further
improve the energy density and safety of the next battery
generation.
665
Quasi-solid and solid-state electrolytes are pro-
mising for solving these issues.
666
Quasi-solid electrolytes are
defined by a liquid electrolyte enclosed in the pores of a solid
matrix (polymer matrix, MOFs, etc.), whereas solid-state elec-
trolytes are solely composed of a solid ionic conductor.
667
The
unique properties of ionic movement in SNPCs have been
recently highlighted as a promising way of improving ionic
conductivity and resolving electrolyte side reactions in quasi-
solid and solid-state electrolytes.
668
The following part reviews
the latest advance of the application of SNPCs in quasi-solid
and solid-state electrolytes.
3.3.1. Quasi-solid-state electrolytes
3.3.1.1. SNPCs in gel polymer electrolytes. Although gel poly-
mer electrolytes (GPEs) present high safety and relatively good
electrochemical properties, the quest to increase the cation
transference number (to achieve high-rate performance and
suppress dendritic growth) and reduce the side reactions at the
electrolyte/electrode interface continues. During the past dec-
ades, nanometre and sub-nanometre polymeric matrices have
been investigated through intensive experimental and theore-
tical practices. For instance, in 2011, Li et al. employed multi-
axis pulsed-field-gradient nuclear magnetic resonance (NMR),
H NMR spectroscopy and synchrotron small-angle X-ray scat-
tering to understand how the PEM structure and morphology
influence the ion transport from the sub-nanometre to micro-
metre scale.
669
Theoretical models such as molecular dynamics
simulations showed that the ionic transport in SNPC polymer
films is ascribed to the negatively charged nanopores (radius of
E0.3 nm).
670
More recently, machine learning was employed to
elucidate the energy barrier descriptors from 126 features, this
approach could reveal key attributes for the design of ion
selectivity sub-nanopores membranes.
671
Coupling cutting-
edge experimental characterisation studies with theoretical
calculations and computing tools could revolutionise the for-
mulation of electrolytes in energy storage systems.
Several research groups demonstrated that Li ions can
penetrate SNPC whereas larger anions and solvent molecules
remain stranded.
392,429
Therefore, employing SNPCs in bat-
teries introduces specific ion selectivity, which leads to a high
Fig. 15 (a) Schematic illustration of ACOF coating with ion selective SNPCs for Li metal anodes. Reproduced from ref. 646 with permission from Wiley-
VCH, Copyright 2021. (b) Illustration of the synthesis of COFs with SNPCs for Li anode coating to achieve stable Li metal anodes. Reproduced from ref.
648 with permission from Wiley-VCH, Copyright 2020. (c) Schematic of the synthesis of COFs with SNPCs on Li metal for Li metal anodes. Reproduced
from ref. 649 with permission from Wiley-VCH, Copyright 2022. (d) Schematic illustration of the ion transport on the Zn metal anode with a zeolite based
protective layer. Reproduced from ref. 657 with permission from Wiley-VCH, Copyright 2023. (e) Schematic illustration of uniform Zn metal plating
facilitated by PSPMA coating with SNPCs. Reproduced from ref. 658 with permission from The Royal Society of Chemistry, Copyright 2023. (f) Schematic
illustration of VRM coating on Zn metal with ion accelerating SNPCs for stable Zn metal anodes. Reproduced from ref. 635 with permission from Elsevier,
Copyright 2022.
Chem Soc Rev Review Article
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2024
cation transference number and provides a physical barrier
limiting the electrolyte decomposition. Depending on the ion
solvation sheath and on the SNPC internal functional groups,
various electrostatic interactions dictate the ion selectivity
and transport. In organic electrolytes, cations and anions
can penetrate the sub-nanochannels with partial solvation
sheath.
430
As the pore size decreases, the number of solvent
molecules removed from the solvation shell increases. Analo-
gically, Lu et al. in situ characterised the dehydration phenom-
ena of alkali metal ions (Li
+
,Na
+
and K
+
) during ion transport in
polymeric SNPCs.
431
Interestingly, they observed partial shred-
ding of water molecules from the hydration sheath, which
allowed the ions to penetrate pores with smaller size than their
solvated size. This fundamental characterization opens exciting
routes for aqueous electrolytes to enlarge the electrochemical
stability window of the electrolyte and suppress water splitting
side reactions. For instance, Li et al. developed a composite
membrane based on 2D vermiculite and 1D CNFs with ZnSO
4
aqueous electrolyte. Based on dielectric relaxation spectroscopy
and molecular dynamics studies, two types of channels form
two corresponding types of confined water molecules, with
different water relaxation processes and water activities. Due
to the fast relaxation process of water in 1D fiber-expanded 2D
capillaries, increasing the CNF ratio to 75% will result in the
maximum Zn
2+
conductivity of 1D/2D membranes, which is far
higher compared to the original 2D vermiculite and 1D CNF
membranes. Meanwhile, the electrochemical activity of con-
fined water restricted by 1D/2D is limited, including a 78%
reduction in corrosion current, and the unique egg-like struc-
ture of the elastic modulus of the pebble-like structure is
further increased by 155% compared to the vermiculite
membrane. Due to the confinement effect of controlled water
and enhanced mechanical strength, non-dendritic zinc electro-
deposition has been achieved, and the permeation of zinc in
the battery is significantly reduced.
432
A different quasi-solid electrolyte formulation approach
involves introducing SNP particles as fillers in gel electrolytes
to promote the cation transference number. Yang et al. engi-
neered a gel polymer electrolyte consisting of a PvDF-HFP
micro-structured matrix and two inorganic fillers (SiO
2
nano-
particles and Zeolite Socony Mobil-5, ZSM-5).
433
As schemati-
cally explained in Fig. 16a, the SiO
2
nanoparticles adsorb PF
6
anions owing to Lewis acid interactions, which favours Li salt
dissociation. Meanwhile, the ZSM-5 particles provide many
sub-nanopores which enhance the ionic conductivity of Li
+
,
thereby efficiently hindering Li dendrite growth (Fig. 16b). To
demonstrate its practical application, the gel polymer electro-
lyte was applied in Li||NMC811 and Li||LFP batteries. The cells
exhibited excellent cycling stability, i.e., 300 cycles with 92%
capacity retention for Li||NMC811 batteries and 500 cycles with
96% capacity retention for Li||LFP batteries. In addition, gel
polymers are also performed as quasi-solid-state electrolyte
for Na batteries. As shown as Fig. 16c, a polyvinylidene
fluoride-hexafluoropropylene copolymer (PH) membrane is
used as a flexible three-dimensional porous polymer matrix
that can incorporate additional microporous polymers.
672
By
introducing super-crosslinked microporous polymers into the
PH host membrane, a continuous polymer network is created
that can capture liquid electrolytes and create ‘‘liquid SNPCs’’.
Compared to modern separators used in liquid electrolyte
systems, these quasi-solid polymer electrolytes offer higher
safety and flexibility without sacrificing the high ion conduc-
tivity of traditional liquid electrolytes. Specifically, we synthe-
sized super-crosslinked microporous polymer membranes
incorporating polyfuran or polypyrrole into the 3D PH struc-
ture. These membranes demonstrate the ability to immobilize
liquid electrolytes within the microporous polymer matrix,
resulting in quasi-solid electrolytes with high ion conductivity,
stability, and cycling life when applied in lithium and sodium
metal batteries. Due to a certain affinity for liquids, electrolytes
are easily trapped in the microporous polymer. Therefore, GPEs
can provide key advantages over SPEs and liquid electrolytes,
including acceptable ion conductivity (410
4
at room tempera-
ture), good wetting of alkali metal anodes, and maintaining
close interface contact with the cathode and anode. More
Table 5 Summary of the application of SNPCs for anode materials in
various battery systems
Materials Categories SNPCs size (nm) Batteries Ref.
HC Carbon o1 PIBs 394
Graphdiyne Carbon 0.55–0.56 LIBs 611
HPCNS Carbon 0.7–1.3 PIBs 612
SiC-CDC Carbon 0.52–1.0 PIBs 395
HC Carbon 1.0 SIBs 613
SC-NSs Carbon 0.45–0.90 SIBs/PIBs 610
Soft carbon Carbon o1 PIBs 396
C-1300 Carbon 0.8 SIBs 397
PC Carbon 0.43–0.73 PIBs 609
HC Carbon o0.5 SIBs 398
CMP Polymer o1 LIBs/SIBs 660
CMP Polymer 0.4–0.6 LIBs/SIBs 400
CMP Polymers 0.7–1.1 LIBs 629
CMP Polymers 0.5–1.5 SIBs 347
CMP Polymers 0.5 PIBs 402
CMP Polymers o1 ABs 628
Ti
1.73
O
41.04
Intercalation 1.15 LIBs/SIBs/PIBs 619
MoS
2
/Gr/C Intercalation 0.98 LIBs 620
DMcT-MoO
3
Intercalation 1.04 SIBs 621
MXene Intercalation 1.02 LIBs 622
HTO-PANI Intercalation 9.3–9.9 SIBs/PIBs 625
SUCNs-SF Conversion 0.6 LIBs 632
Graphene/MoS
2
Conversion o1 LIBs 661
MoS
2
Conversion o1 LIBs 633
Ni
2
P@C-N Conversion 1.0 SIBs 634
Si@ZIF-8 Alloying B1.1 LIBs 403
ZIF-8-C@PP Alloying 0.5–1.0 LIBs 404
P@N-MPC Alloying o1 SIBs 405
HPCNS/P Alloying o1 SIBs 406
N-CBCNT@rP Alloying 0.9 LIBs 407
Cu-OMC@RP Alloying o1 LIBs 408
CMP-Li Li metal 0.5–0.6 LMBs 139
ACOF-coated Li Li metal 0.6 LMBs 646
COF-Li Li metal 0.6 LMBs 648
COF-F6@Li Li metal 0.93 LMBs 662
Ploymer-Li Li metal 0.6–1.2 LMBs 12
BDLC Li Metal 0.68 LMBs 652
UiO-66-Li Li metal 1.0–1.2 LMBs 149
ZnA@Zn Zn metal 0.35 AZBs 657
VRM@Zn Zn metal 0.4–1.2 AZBs 663
Zn@MGs Zn metal 0.92 AZBs 664
Review Article Chem Soc Rev
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
This journal is © The Royal Society of Chemistry 2024 Chem. Soc. Rev.
importantly, microporous polymer candidate materials with
small pore sizes and high porosity can absorb large amounts
of liquid electrolyte while providing minimal electrolyte leakage
and maintaining the ability to suppress the formation and
growth of dendritic substances due to their small pore size.
Because of this, the method of trapping liquid electrolytes in
microporous polymers has been widely used as gel polymer.
Full-cell lithium and sodium batteries containing these micro-
porous polymer electrolyte membranes exhibit comparable rate
and interface performance to traditional liquid electrolytes, but
with significantly improved cycling performance and coulom-
bic efficiency. Furthermore, the polymer of inherent micropor-
osity (PIMs) has been widely used in functional membranes in
recent years because of the stable structure micropores formed
by inefficient molecular stacking. Thus, GPE was prepared
using bacterial cellulose (BC) as a support membrane and
partially acylated PIM-1 (named PIM-CONH
2
) as a functional
layer for use in lithium metal batteries (LMBs), as shown in
Fig. 16d. Due to the presence of amide groups, the formed PIM-
CONH
2
/BC GPE exhibits a high Li
+
transport number (0.76),
good cycling performance (retains 83% capacity after 200 cycles
at 0.5C), and good rate capability in LiFePO
4
|GPE|Li batteries.
The layered porous structure in PIM-CONH
2
/BC GPE can reg-
ulate the Li
+
transport behavior, and the presence of amide
groups promotes the uniform dissolution/deposition of Li
+
(Fig. 16e). Acylated PIMs in GPE can play an important role in
improving interface affinity and suppressing lithium dendrite
growth. Similarity, Zou et al. employed bacterial cellulose (BC)
as the support membrane and partially acylated PIM-1 (named
PIM-CONH
2
) as the functional layer, which was performed as
the electrolyte membrane of lithium metal batteries (LMBs).
673
The PIM-CONH
2
/BC GPE, designed using the vascular system
as a basis, had a multilayer porous structure. This structure
enhanced charge transport and reduced volume changes dur-
ing the cycling process. The amide group in GPE acted as the
main transport site. It regulated the nucleation mode of
lithium, induced uniform deposition of Li, and inhibited the
growth of lithium dendrites. This study provides a new strategy
for improving the safety and stability of LMBs to meet the
demand for functionalized LMB electrolytes.
3.3.1.2. SNPCs in MOF quasi-solid electrolytes. Solid-state
batteries are gaining attention as a potential energy storage
system due to their high energy density and improved safety.
However, the high resistance between the solid-state electrolyte
and electrode leads to slow cation transport. MOF formulation
is adaptable to create high specific area SNPC structures, which
can accommodate ions and small molecules. Sub-nano con-
fined ions/solvent molecules exhibit chemical and physical
properties that are dramatically altered from those in bulk
electrolytes. The physical confinement and chemical interac-
tions between the ions/solvent molecules with metal sites
inside the MOF nano-channels can facilitate the ion transport,
regulate the SEI formation, increase the temperature decom-
position of the electrolyte and adjust the electrolyte up-
take.
18,434,435
Inspired by this, a quasi-solid-state electrolyte
(QSSE) based on a solvate ionic liquid (SIL) confined in nano-
cages of UIO-66 (SIL/UIO-66) is developed. The SIL/UIO-66 QSSE
benefits from the spatial confinement of TFSI
by the UIO-66
pores and the strong chemical interactions between SIL and
metal atoms. Consequently, the SIL/UIO-66 QSSE exhibits high
ionic conductivity and compatibility with electrodes. As a
Fig. 16 (a) Schematic illustration of SZ-GPE. (b) Current–time response of SZ-GPE in Li||Li symmetric cells after applying a constant potential (10 mV).
Inset: Nyquist plots before and after polarization and corresponding fitting curves. Reproduced with permission, from ref. 433 Copyright 2022 c. (c)
Schematic representation of Li or Na plating/stripping on traditional liquid/separator-based (CG or GF) cells compared to cells based on advanced Li-
HCFu-PH or Na-HCPy-PH electrolyte membranes. Reproduced with permission, from ref. 672 Copyright 2020 John Wiley and Sons. (d) Schematic
formation of the PIM-CONH
2
/BC composite GPE. (e) Mechanism of lithium dendrite growth inhibition by PIM-CONH
2
/BC GPE. Reproduced with
permission from ref. 673 Copyright 2023 Elsevier.
Chem Soc Rev Review Article
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2024
result, Li|QSSE|LFP cells demonstrate excellent rate capability
and cycle stability in a wide temperature range of 25–90 1C. This
study presents a practical approach for fabricating safe solid
electrolytes with excellent compatibility and long cycle life for
high-performance QSSE LIBs.
Chang et al. prepared a safe quasi-solid electrolyte by con-
fining a Li-based organic liquid electrolyte inside the SNPCs of
a MOF.
435
The as-prepared quasi-solid electrolyte exhibits high
safety (non-flammability) and a large electrochemical stability
window (5.4 V vs. Li/Li
+
), owing to electrolyte confinement; the
Li solvation sheath became more compact and therefore more
energy is needed for the solvated PC molecules to be oxidized
(Fig. 17a). Meanwhile, the energy density was also improved
since only a small amount of liquid electrolyte was necessary to
fill the MOF sub-nanochannels. The Li||NMC811 pouch cells
cycled for more than 300 cycles at room temperature and at
90 1C with an excellent Coulombic efficiency even though the
cathode loading was 20 mg cm
2
and electrolyte up-take was
0.23 mLcm
2
. Similarly, ionic liquids, which are considered as a
safer alternative to organic solvent, have been infiltrated in the
pores of a self-assembled zeolite imidazole framework (ZIF-8)
3D structure (Fig. 17b).
436
Recently, a MOF was decorated with
halogen atoms inside the channels, leading to a high ionic
conductivity of 2.16 10
4
Scm
1
at room temperature, a large
electrochemical stability window (4.8 V) and dendrite
suppression.
437
Besides, the MOF nanocrystals were continu-
ously grown on polyimide fibres. The SNPCs of the MOF were
infiltrated with a Li-containing ionic liquid electrolyte; mean-
while, the SNPCs between the interconnected fibres were filled
in with the PvDF–LiTFSI polymer composite. The SNPCs of the
Fig. 17 (a) Schematic illustration of the porous PSS polymer with CuBTC MOF decorated inside its channels and its advantages. Reproduced with
permission from ref. 435 Copyright 2022 Springer Nature Limited. (b) Schematic illustration of the hierarchically self-assembled MOF network as a 3Dion
conductor with continuous Li
+
transport. Reproduced with permission from ref. 436 Copyright 2022 John Wiley and Sons. (c) The preparation of SIL/
UIO-66 and the assembling of quasi-solid-state batteries. Reproduced with permission from ref. 459 Copyright 2022 John Wiley and Sons. (d) MOF-
based SSI formed on LMA with immobilized anions and ionic channels for fast Li
+
transport. Reproduced with permission from ref. 149 Copyright 2020
Elsevier. (e) Schematic illustration of MOF–PVDF GPE with anions immobilized for lithium–sulfur batteries. Reproduced with permission from ref. 674
Copyright 2017 American Chemical Society. (f) Illustration of the assembly and design of solid-state batteries based on UN-LiM-IL Sea. Reproduced with
permission from ref. 675 Copyright 2023 American Chemical Society.
Review Article Chem Soc Rev
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
This journal is © The Royal Society of Chemistry 2024 Chem. Soc. Rev.
MOF network provided high Li
+
conductivity owing to the
spatial anion confinement in the sub-nanopores and the
Lewis acid interactions between anion-metal sites. Such a
quasi-solid electrolyte presented a high ionic conductivity of
4.08 10
4
Scm
1
at 30 1C and enhanced the cycling
performance of Li||LFP batteries. Furthermore, the MOF struc-
ture can be tuned by altering the ligands or the metal sites. In
addition, solid-state electrolytes are widely used in solid-state
lithium batteries to help improve the slow transmission
kinetics caused by high interfacial resistance between the solid
electrolyte and the electrode. Thus, a quasi-solid-state electro-
lyte (QSSE) based on solvent ionic liquid (SIL) confined in
UIO-66 nanocages is prepared.
459
Due to the effective spatial
confinement of TFSI
through UIO-66 SNPCs and the strong
chemical interaction between SIL and metal atoms, SIL/UIO-66
QSSE exhibits high ion conductivity and good electrode com-
patibility (Fig. 17c). Therefore, Li|QSSE|LFP batteries demon-
strate excellent rate performance and cycling stability in a wide
temperature range of 25–90 1C. This study provides a practical
strategy for preparing safe solid electrolytes for high-
performance QSSE lithium-ion batteries. Apart from improve-
ment in ionic conductivity, the high interfacial resistance
between electrolyte and electrolyte is also a crucial factor to
suppress cations transportation. A QSSE based on a solvent-
type ionic liquid (SIL) confined in the UIO-66 nanocage was
prepared (Fig. 17d).
149
Due to the effective space restriction of
TFSI- in the UIO-66 pore and the strong chemical interaction
between SIL and metal atoms, SIL/UIO-66 QSSE exhibits low
interfacial resistance and good electrode compatibility. There-
fore, the Li|QSSE|LFP battery shows excellent rate performance
and cycle stability over a wide temperature range of 25–90 1C.
Additionally, a universal interphase design protocol for durable
lithium metal anodes has been proposed by Wu’s group. It is
recommended to cycle high-capacity Li metal anodes using ion
conductors with high lithium transport numbers and high
diffusion coefficients. An anion-binding semi-solid interphase
constructed with metal–organic frameworks as an ion transport
rectifying layer has been used to meet the requirement of
suppressing Li dendrite growth on practical Li metal anodes.
By using protected thin Li metal, high-loading cathodes, and
low electrolyte usage, high energy density Li metal batteries
have been demonstrated.
S/Se cathode batteries are also attracting great attention for
promising next-generation rechargeable energy storage sys-
tems. They offer advantages like traditional liquid metal oxide
batteries and address the drawback of liquid-based metal-
chalcogen batteries, which is the ‘‘shuttle effect’’. However,
the poor compatibility of the electrode/electrolyte interface and
the low ion conductivity of solid-state electrolytes are key
issues hindering their practicality. Thus, a gel polymer electro-
lyte (GPE) modified with a metal–organic-framework (MOF) is
utilized to stabilize the lithium anode in a Li–S battery
(Fig. 17e).
674
The MOF skeleton contains abundant pores,
which immobilize large-size polysulfide anions and cage elec-
trolyte anions, resulting in a uniform flux of Li
+
and homo-
geneous Li deposition. In addition, Wu et al. have designed
MOFs by incorporating two types of ionic liquids (ILs) for
creating quasi-solid electrolytes.
675
The resulting MOF-IL elec-
trolytes provide uninterrupted ion transport channels with
functional sulfonic acid groups that serve as lithium ion hop-
ping sites, thus enhancing Li
+
transport in both the bulk
and at the interfaces (Fig. 17f). These quasi-solid MOF-IL
electrolytes exhibit competitive ionic conductivities of over
3.0 10
4
Scm
1
at room temperature, wide electrochemical
windows over 5.2 V, and good interfacial compatibility. Apart
from employing electrolytes, MOFs can also be used as an
electrolyte additive to effectively suppress the generation of
side reactions. For example, when Li
6
PS
5
Cl (LPSCl) is used as
an electrolyte, it is easy to release hydrogen sulfide (H
2
S) in a
humid environment, which affects the performance of bat-
teries. To solve this problem, the Jin group used ZIF-8 as a
desiccant in LPSCl without changing the Li
6
PS
5
Cl electrolyte or
electrode structure. By introducing highly ordered porous
materials, we have demonstrated the stable cycling perfor-
mance of the configured LPSCl SSEs battery, due to the effective
and lasting desiccating effect.
676
Porous organic frameworks (POFs) are porous materials
constructed from organic matrices. The inherent pores of POFs
meet the demand for constructing rapid Li
+
diffusion SNPCs to
achieve high ion conductivity in solid polymer electrolytes
(SPEs). Although some studies have reported SPEs on POF
substrates, most are limited to thin-film frameworks of func-
tional groups (such as boronic esters, b-ketoimines, and tria-
zines), and the transport capacity of Li
+
is restricted. In
addition, stable covalent bonds are necessary to construct
porous materials with rich porous structures to ensure the
stability of fast Li
+
transport channels in SPEs. Porous aromatic
frameworks (PAFs) are a new type of POFs with rigid skeletons,
open structures, high specific surface areas, and excellent
stability, composed of irreversible carbon–carbon bonds. The
abundant SNPCs of PAFs allow for free diffusion of Li
+
. There-
fore, constructing imidazole anion-type PAFs enables easy
dissociation and efficient transport of Li
+
in PAF-based SPEs.
To ensure high Li
+
transport in SNPCs, a high Li
+
content is
necessary according to the classical equation: s(T)=Pn
i
q
i
m
i
.
The Sonogashira–Hagihara coupling reaction was applied to
prepare a new imidazole-based PAF (PAF-220), which was then
lithiated to obtain a highly lithium-ion conductive lithium-rich
anionic PAF (PAF-220-Li) (Fig. 18a).
677
The binding energy
between imidazole anions and Li
+
is low, which facilitates the
dissociation of Li
+
. The conjugated structure of benzimidazole
can delocalize negative charges and further reduce the binding
energy between imidazole groups and Li
+
. In addition, the
anionic framework can restrict the movement of anions, while
the abundant channels in the single-ion PAF-220-Li only allow
Li
+
diffusion (Fig. 18b). PAF-220-Li was further vacuum-injected
with LiTFSI and combined with (polyvinylidene fluoride-co-
hexafluoropropylene, PVDF-HFP) to obtain a quasi-solid-state
polymer electrolyte (PAF-220-QSPE) with an ion conductivity of
0.206 mS cm
1
at room temperature, up to 0.76. It is worth
noting that the rigid structure of PAF-220-Li grants it good
stability, providing excellent physicochemical stability with a
Chem Soc Rev Review Article
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2024
wide electrochemical window (5.0 V) for PAF-220-ASPE. There-
fore, this work provides a valuable strategy for high-
performance SPE based on functional PAF and may become a
practical means for the design of solid-state LIBs. Similarity,
Wang’s group reported a strategy to confine low molecular
weight polyethylene glycol (PEG, molecular weight = 800) into
the channels of COFs, a solvent-free strategy is used to effi-
ciently promote Li
+
ion movement within the SNPCs of COFs,
enabling rapid and stable conduction of cations over a wide
temperature range (Fig. 18c and d).
678
SNPCs applied in quasi-solid-state electrolytes provide
many benefits. The SNPCs can be tailored to achieve specific
properties depending on the application, which can increase
the cation transference number by allowing cation conduction
while hindering anion movement. This property is critical in
metal batteries to mitigate dendrite growth in metal anodes
and improve the cycling rate performance for metal-ion bat-
teries. Nonetheless, fully understanding the local properties of
ions inside the SNPCs as well as at the interface of bulk
electrolyte/SNPCs remains challenging.
3.3.2. Solid-state electrolytes
3.3.2.1. SNPCs in solid polymer electrolytes. Solid-state poly-
mer electrolytes (SPEs) have garnered significant interest due to
their exceptional properties in lithium-ion batteries, including
Fig. 18 (a) Schematic representation of the preparation of PAF-220-SPE and PAF-220-ASPE. (b) Illustrative drawings showing the Li
+
transport
mechanism in PAF-220-SPE and PAF-220-ASPE. Reproduced with permission from ref. 677 Copyright 2023 John Wiley and Sons. (c) Schematic
illustrations of Li
+
transport in COFs and (d) structural representations of CD-COF, COF-5, COF-300, and EB-COF. Reproduced with permission from ref.
678 Copyright 2019 American Chemical Society.
Review Article Chem Soc Rev
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
This journal is © The Royal Society of Chemistry 2024 Chem. Soc. Rev.
low flammability, resistance to leakage, superior thermal sta-
bility, excellent processability, high flexibility, and enhanced
safety levels. Polyethylene oxide (PEO) is commonly reported as
SPEs since it was discovered that Li ions can hop from one
ethylene oxide (EO) group to the next with polymer chain
motion, thereby demonstrating some ion conduction.
679,680
By definition, in crystalline PEO, the molecular chains are
static, thereby hindering ionic conduction. Therefore, PEO-
based solid polymer electrolytes suffer from extremely low
ionic conductivity at room temperature due to the crystallinity
of the PEO chains and can only be applied as electrolyte above
their temperature (application above B60 1C). Contrarily,
Bruce et al. reported Li conduction in a highly ordered crystal-
lized PEO
6
:LiAsF
6
system.
681
According to X-ray diffraction
(XRD) analysis and ab initio simulations, they demonstrated
that in the EO : Li ratio 6: 1, the PEO chains adopt a double
helical structure forming a cylinder shape.
438
The Li ions are
located inside the cylinder whereas the anions remain outside.
This structure acts as SNPCs with the distance between Li
and EO units ranging from 2 to 2.3 Å. In order to increase
the ionic conductivity of solid polymer electrolytes, another
polymer was explored by Qi and Hu’s groups.
439
They demon-
strated the intercalation of Li ions in elementary cellulose
nanofibrils (CNFs). The copper (Cu) coordination with the
CNFs allowed the expansion of the interspacing between mole-
cular chains from 0.39 nm to 0.87 nm (Fig. 19a and b).
The decoupling of Li ions resulted in excellent ionic conductiv-
ity and transference number (Fig. 19c) and cost-efficiency,
scalability, and flexibility (Fig. 19d). To prove the practical
application of the as-prepared solid polymer electrolyte,
an all-solid-state pouch cell based on Li metal anodes and
LFP cathodes demonstrated exceptional cycling performance
with a capacity retention of 94% after 200 cycles at room
temperature (Fig. 19e and f). Furthermore, other ions (Na
+
and Zn
2+
) and polymer (chitosan, CMC, alignate, XG/CNF)
combinations are also suitable for creating analogous solid
polymer electrolytes.
Another approach to improve the ionic conductivity of solid
polymer electrolytes through SNC engineering involved utiliz-
ing MOFs grown on polymer networks. Zeng et al. in situ grew a
MOF functionalized with high-density electronegative groups
(Zr-BPDC-2SO
3
H) on bacterial cellulose (Fig. 19g and h).
440
The
in situ growth enabled minimization of the interfaces between
MOF particles and provided a continuous ion conduction
pathway. They demonstrated a Li transference number of 0.88
(Fig. 19i), achieving a high specific capacity of 119 mA h g
1
at a
high cycling rate (3C). Recently, Hu et al. employed a layered
composite solid polymer electrolyte composed of Uio66-NH
2
MOF particles and the PVDF-HFP-LiTFSI-SN-FEC solid
polymer.
441
The MOF layer acted as an anion sieve owing to
the sub-nanopores of the MOF particles,while facilitating high
Li conduction. Therefore, the composite solid electrolyte exhib-
ited an ionic conductivity of 0.245 mS cm
1
. The Li ions forced
through the sub-nanopores of the MOF layer generated a uni-
form Li flux, which benefit excellent Li plating and stripping
cycling performance.
3.3.2.2. SNPCs in inorganic solid electrolytes. Inorganic solid
electrolytes can be categorized into three large families accord-
ing to their chemical structure:
682,683
oxides, which regroups
perovskite conductors (ABO
3
–type, the general formula is
Li
x
La
2/3x/3
TiO
3
, LLTO), anit-perovskite conductors (e.g.,Li
3
OCl
and Li
3
OBr), NASICON conductors (Na
1+x
Zr
2
P
3x
Si
x
O
12
and
LiM
2
(PO
4
)
3
(M = Zr, Ti, Hf, Ge or Sn)) and garnet-type con-
ductors (e.g.,Li
5
La
3
M
2
O
12
(M = Nb, Ta), orthosilicate garnets
(A
3II
B
2III
(SiO
4
)
3
), A
3
B
5
O
12
(A = Ca, Mg, Y or Ln = La, or rare-earth
elements; B= Al, Fe, Ga, Ge, Mn, Ni, and V), cubic Li
7
La
3
Zr
2
O
12
(LLZO)); sulfide-type, which reassembles thio-LISICONs
with the formula Li
4x
M
1y
M0
y
S
4
(M = Si, Ge, and M’ =P, Al,
Zn, and Ga), LGPS family (Li
10
GeP
2
S
12
,Li
10
SnP
2
S
12
, and
Li
9.54
Si
1.74
P
1.44
S
11.7
Cl
0.3
), arygodrites-type with the common
formula Li
6
PS
5
X (X =Cl, Br, and I); halide-type with the for-
mula—Li
a
MX
b
, where X represents I, Br, Cl, and F while M is a
metal element (e.g.,Li
3
InCl
6
,Li
3
YCl
6
,Li
3
ErCl
6
, and Li
2
ZrCl
6
).
Recently, Kang et al. reported the control of superionic
conduction in ternary halides (e.g.,Li
3
MCl
6
[where the metal
(M) is Y or Er]) by manipulating the in-plane lithium diffusion
pathways and the stacking layer distances. These two factors
are negatively correlated with each other through the partial
occupancy of M, which acts as both a diffusion inhibitor and a
pillar to maintain interlayer distances. The authors demon-
strated the critical range or ordering of M in ternary halides and
showed the achievement of high ion conductivity by adjusting
the simple M ratio. This work provides a general design
standard for superionic ternary halide electrolytes.
684
Besides,
the inorganic solid electrolytes contain metal cations, which
take part in electrochemical reactions. As shown in Fig. 20a, the
Li
+
conduction occurs through different mechanisms: intersti-
tial hopping, interstitial knock-off and vacancy hopping.
666,685
The atomic sites in inorganic solid state electrolytes can be
compared to sub-nanopores and the cation conduction analo-
gously occurs through sub-nanochannels. Kato et al. discovered
a superionic conductor Li
9.54
Si
1.74
P
1.44
S
11.7
Cl
0.3
, which showed
the highest Li
+
conduction reported for inorganic solid-state
electrolytes (Fig. 20d).
442
They demonstrated that the double
substitution with aliovalent-ion doping (Si and Cl atoms)
creates a 3D Li
+
conduction pathway in Li
9.54
Si
1.74
P
1.44
S
11.7
Cl
0.3
.
The 1D pathway is characteristic of the LGPS family, while the
2D pathway could be attributed to the small amount of Cl
atoms located in the Cl(1) (8g) sites situated in P(2b)X
4
tetra-
hedra (Fig. 20d and e). Therefore, the ionic conductivity in
solid-state electrolytes is closely related with the conduction
pathways. Inorganic solid electrolytes are polycrystalline, i.e.,
contain grains and grain boundaries. The ionic conductivity in
the grains, across the grain boundaries and in the grain
boundaries varies largely depending on the crystallographic
structure of the inorganic solid electrolyte. This leads to two
competing conduction pathways: from one grain to the neigh-
bouring grain, i.e., across the grain boundary, named ‘‘the
granular pathway’’, and within the grain boundaries, named
‘‘the grain boundary pathway’’ (Fig. 20b and c). The granular
pathway will be preferential when the grain boundaries are
more resistive than the grain, for instance, in LATP and Li
3
OCl.
Chem Soc Rev Review Article
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2024
Otherwise, the grain boundary pathway will dominate, which is
the case for sulfide-based electrolyte and some oxides such as
LLZO.
443
Several strategies have been explored to increase the ionic
conductivity in inorganic solid-state electrolytes. Although not
fully understood yet, it was demonstrated that employing a
doping and/or substitution strategy can dramatically increase
the ionic conductivity. For instance, the introduction of 0.3 Al
mol per unit in the LTP crystal enriches the M3 site in Li atoms,
thereby opening a new conduction channels M1-M3-M1 site
with an activation energy between M1/M3 and M3/M1 much
lower than that between M1/M1.
444
The new conduction chan-
nels enable LATP to exhibit a much higher ionic conductivity
than the parent LTP crystal. However, substituting Ge
4+
by Si
4+
or Sn
4+
in LGPS-type solid electrolytes resulted in lower ionic
conductivities. It is thought that Si doping narrows the sub-
nano channels for Li
+
migration; further fundamental investi-
gations on Li
+
conduction pathways at a sub-nanoscale are
necessary to develop novel superionic conductors.
683
The
substituted and/or doped element intrinsic characteristics
Fig. 19 (a) Schematic illustration of the structure of CNFs. Bottom: Coordination of Cu
2+
ions with the hydroxyl groups of cellulose which opens the
spacing between the molecular chains, creating cellulose molecular Li
+
-conducting channels in the CNFs. (b) Scanning electron microscopy (SEM)
image of the CNFs. (c) Transference number is plotted against Li
+
ionic conductivity for Li–Cu–CNF and for other SPEs PEO-inorganic composites,
cross-linked polymers, and a high-Li-concentration electrolyte. (d) Digital photograph of a 1-m-long roll of Li–Cu–CNF membrane. (e) Illustrationofa
Li|Li-Cu-CNF|LFP all solid-state battery Li-metal as anode, Li–Cu–CNF paper as electrolyte and LFP/CNT/Li–Cu–CNF as a cathode. The Li–Cu–CNF
(green fibres) enables transport of Li
+
ions (yellow arrows), and CNTs (red fibres) in the cathode enable electron transport (red arrow). (f) Cycle
performance of a solid-state LiFePO
4
cell made using Li–Cu–CNF ion-conducting binders in the cathode, Li–Cu–CNF electrolyte, and a Li-metal anode.
Reproduced with permission from ref. 439 Copyright 2023. Springer Nature. (g) Structure of SSBs with flexible cross-linked MOF chains as SE and the
crystal structure of Zr-BPDC-2SO
3
H. Color code: Zr, purple; O, blue; C, gray; and S, yellow. (h) Bottom–up synthesis of cross-linked MOF chains on the
BC skeleton. (i) The ionic conductivities of Zr-BPDC-2SO
3
HforLi
+
,Na
+
,K
+
, and Zn
2+
. Reproduced with permission from ref. 440 Copyright 2021
American Chemical Society.
Review Article Chem Soc Rev
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
This journal is © The Royal Society of Chemistry 2024 Chem. Soc. Rev.
(e.g., atoms size, valence, etc.) greatly influence greatly the bulk
ionic conductivity. Besides, the crystal phase also seriously
impacts the ionic conductivity due to the different arrangement
of atoms. For instance, LLZO solid-state electrolyte has two
crystal structures, the tetragonal phase (t-LLZO) and cubic
phase (c-LLZO). In t-LLZO, all the Li
+
positions in the crystal
structure are occupied, whereas in c-LLZO, one site is occupied
while the two neighbouring sites are vacant. The c-LLZZO
exhibits approximatively twice the ionic conductivity of t-
LLZO.
686
Another challenge suppressing the development of all-solid-
state batteries is the high interface resistance between the
electrolyte and the electrode, which has a significant impact
on the overall performance of the battery.
687
In all-solid-state
batteries, electrochemical reactions occur at the solid–solid
electrolyte–electrode interface. Ions diffuse from the electrolyte
to the electrode through their interconnected region and
undergo oxidation–reduction reaction at the electrolyte–
electrode interface in contact with the active material and
electrons. Therefore, maintaining a good solid–solid electro-
lyte–electrode interface in the battery is crucial for achieving
stable charge transfer reactions. One of the main causes of high
interface resistance is the chemical and electrochemical
instability between the solid electrolyte and the electrode, such
as the space charge layer of local ion vacancies in the electrolyte
near the electrode, which severely limits the rate capability of
the battery due to the chemical potential difference between the
electrode and the electrolyte, causing ions to be extracted from
the electrolyte to the electrode side.
688
In addition, poor contact
between the solid electrolyte (especially inorganic solid electro-
lyte) and the electrode is also a major issue affecting ion
diffusion.
689
The volume changes of cathodic and anodic
materials during charge and discharge processes may lead to
the loss of effective contact between the electrode and the solid
electrolyte, thereby limiting the conduction of ions in the
interface area. Furthermore, it is worth noting that the diffu-
sion layer of elements formed at the electrolyte–electrode inter-
face is also an important resistance for the deterioration of
interface stability during the cycling process.
690
Since the solid
electrolyte is physically inflexible, the battery manufacturing
process usually requires additional heating steps to improve
the adhesion between the electrode and the electrolyte (oxides).
Therefore, an element mutual diffusion region forms at the
solid–solid interface, accompanied by the formation of signifi-
cant and inevitable interface resistance.
Various methods have been proposed to improve the inter-
face between the solid-state battery electrolyte and electrode.
Interface resistance is mainly attributed to the incompatibility
between the solid-state electrolyte and electrode, due to effects
like space charge, chemical and electrochemical instability, and
interdiffusion. Coating ion-conductive and electron-insulating
buffer layers on the electrode is an effective way to reduce
Fig. 20 (a) Schematic of the Li-ion transport in solid inorganic electrolytes. Reproduced with permission from ref. 666 Copyright 2020. Springer Nature.
(b) Schematic of the granular (top) and GB (bottom) conduction in polycrystalline materials. Reproduced with permission from ref. 443 Copyright 2017.
American Chemical Society. (c) Arrhenius conductivity plots for the LGPS family and Li
9.54
Si
1.74
P
1.44
S
11.7
Cl
0.3
. (d) Crystal structure of
Li
9.54
Si
1.74
P
1.44
S
11.7
Cl
0.3
. The framework structure consists of 1D polyhedral chains. (e) Nuclear distributions of Li atoms in Li
9.54
Si
1.74
P
1.44
S
11.7
Cl
0.3
at
25 1C. Reproduced with permission from ref. 442 Copyright 2016. Springer Nature.
Chem Soc Rev Review Article
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2024
incompatibility. Ohta et al. found that coating Li
0.33
La
0.557
TiO
3
(LTO) on LiCoO
2
(LCO) particles significantly improved the rate
performance of all-solid-state lithium metal batteries based on
sulfide electrolytes, due to reduction in the space charge layer
effect.
691
Besides, introducing an artificial SEI layer with elec-
tron insulation and ion conduction is also an effective method
to reduce interface instability. Goodenough’s team illustrated
the formation of an ion-conductive passivation layer composed
of Li
3
P and Li
8
ZrO
6
on LiZr
2
(PO
4
)
3
solid electrolyte, in contact
with Li metal.
692
In addition, enhancing the electrode–electro-
lyte contact area through electrode–electrolyte nanocomposites
is a commonly employed method, using techniques such as
ball milling, pulsed laser deposition (PLD), and softening glass
electrolytes. Hayashi et al. reported that a significant improve-
ment in the solid–solid interface contact area between the
electrode and electrolyte was achieved by obtaining electrode–
electrolyte nanocomposites through mechanochemical balling.
The large volume changes in the electrode during charge and
discharge processes would lead to significant interface stress
changes, posing a serious problem. Decorating an interface
layer that can closely contact the electrode is an interesting
approach. Yamamoto et al. reported that introducing an NbO
2
layer at the LCO electrode–solid electrolyte interface effectively
mitigated interface stress during delithiation.
693
Solid-state electrolytes are showing increasingly attractive
prospects and potential. Scientists and engineers still need to
work hard to develop solid-state electrolytes that are commer-
cially competitive in terms of ion conductivity, interface impe-
dance, mechanical strength, and electrode compatibility, while
also being cost-effective. It should be noted that different
advantages should be effectively utilized according to different
applications, such as high-power density for portable electronic
products and electric vehicles, and low maintenance costs for
smart grid energy storage, in addition to the required high
energy density. Exploring new technologies, such as optimizing
battery design structures to improve overall performance, is
also essential. Furthermore, modifying existing battery manu-
facturing processes or reinventing new manufacturing technol-
ogies for all-solid-state lithium metal batteries is also important
for short-term practical applications.
3.4. SNPCs in separators/interlayers
For rechargeable batteries, the separator plays a crucial role,
whose main function is to separate the cathode and anode
electrodes to prevent short-circuiting while allowing for the
rapid transfer of the ionic charge carrier required by the circuit
as the current passes through the cell.
694,695
They should be
good electronic insulators and could conduct ions through
intrinsic ionic conductors or by soaking electrolytes. They
should minimize any processes that adversely affect the elec-
trochemical energy efficiency of the battery.
For example, the separator is a very crucial component that
influences the mass transport of cations across the electrodes,
which affects the anode dendrite growth.
435,696
However, the
separator currently in use is prone to cause an in-homogenous
cation flowing pattern, leading to nonuniform cation transfer
that renders dendrite growth.
697
To address this issue, to
design a separator with appropriate and uniformly distributed
SNPCs may aid in the propagation of anode dendrites. For
instance, Rajendran et al. discovered a carbon nanomembrane
(CNM) with an incredibly thin thickness of 1.2 nm (Fig. 21a).
159
This remarkable CNM contained SNPCs measuring 0.7 nm,
which acted as an interlayer to control the movement of Li
+
during mass transport. The presence of these SNPCs improved
the contact and wetting properties at the electrode and electro-
lyte interface, promoting rapid ion diffusion. Compared to
using a Celgard separator alone, the Li-ion transference num-
ber increased to 0.67 when the CNM/Celgard separator was
employed. A higher Li
+
transference number effectively delayed
dendrite nucleation, resulting in nearly stable cycling for more
than 600 hours. This extended cycling time demonstrated
successful dendrite suppression, in contrast to the control
experiment where Li dendrites were observed. Additionally,
the CNM/Celgard separator exhibited exceptional mechanical
strength (Young’s modulus B10 GPa), preventing dendrite
penetration. Besides, the materials with low mechanical
strength restrain their application as separator in batteries.
The larger and wider pore size distribution of the separator can
create steep local concentration gradients at the electrode–
electrolyte interface, which further increases the uneven
deposition of ions and the formation of fibrous, mossy, and
dendritic anode metals.
698,699
Recently, in-depth research on
separator modification has been implemented. A separator
with high mechanical strength can prevent dendrite penetra-
tion. A separator with SNPCs can regulate the uniform ion flux
near the electrode surface, thereby promoting uniform metal
nucleation and growth. Archer and colleagues developed a porous
separator with high mechanical moduli and easy ion transport.
It comprises a well-ordered porous Al
2
O
3
sheet between micro-
porous poly(vinylidene fluoride-co-hexafluoropropylene) (PVDF-
HFP) polymer layers, forming a sandwich-type composite
architecture. The physical properties, including the elastic mod-
ulus and ionic conductivity, were influenced by the pore size,
which is crucial for electrochemical performance.
700
The elastic
modulus of the PVDF-HFP/Al
2
O
3
separator is approximately
0.4 GPa. The ionic conductivity of the PVDF-HFP/Al
2
O
3
separators
was measured at 1 10
3
Scm
1
. Following optimization, the
separators utilizing Al
2
O
3
demonstrated superior performance in
terms of low internal resistance, high CE, and long cycling
performance. Additionally, MOFs and COFs with a highly uniform
SNPC structure make them a promising candidate for construct-
ing separators. Qi et al. prepared a Ti-MOF modified polypropy-
lene/polyethylene separator for rechargeable batteries. Thanks to
the better electrolyte wettability, less resistance, highly intrinsic
SNPCs and Lewis acid characteristics of Ti-MOFs, this modified
separator exhibited the smooth transport of Li
+
.
522
Similarity, a
thin film of covalent organic frameworks (COFs) with azobenzene
side groups branching from the pore walls was employed as a
separator,whichactsasion-hoppingsitesthuspromotingNa
+
migration.
519
For metal–chalcogen batteries, SNPC modified membrane
separators are not only beneficial for cation transport, but also
Review Article Chem Soc Rev
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
This journal is © The Royal Society of Chemistry 2024 Chem. Soc. Rev.
exhibit blocking capability for poly-chalcogen intermediates.
On the other side, excessively small pore sizes may hinder the
diffusion of cations, resulting in retarded redox kinetics. There-
fore, precisely adjusting the size of SNPCs is the key factor for
improving the performance of metal-chalcogen batteries. The
DFT calculations indicate that the SNPCs with a smaller pore
size always exhibit a narrow band gap, which could efficiently
block polysulfide shuttling.
701
Sun’s team synthesized two-
dimensional polymer nanosheets, and then deposited the
modified layers on a separator to form a composite membrane
rich in phenol and triazole units, as well as lithium sites in the
ordered channels of the polymer nanosheets.
702
Due to its
unique SNPCs, Li-COF-based membrane separators can greatly
facilitate the transport of lithium ions while limiting the
diffusion of polysulfides. As a result, the cells obtained showed
higher specific volumes at different rates. Another example
involves the synthesis of a porous SNPC aromatic framework
(SAF) modified separator, which was designed for all-climate
operation of Li–S batteries.
703
After rational molecular engi-
neering design, we developed fully conjugated SAF-3 with
SNPCs (0.97 nm) and the polysulfide migration barrier up to
33.21 eV (Fig. 21b). In addition, Bader charge calculations
(Fig. 21c and d) revealed that SAF-3 transferred more electrons
when interacting with Li
2
S
6
.In situ X-ray photoelectron spectro-
scopy (XPS) and in situ XRD tests show that the SA3-modified
separators can inhibit the transfer of polysulfide and promote
Fig. 21 (a) Schematic representation of the synthesis and transfer process of the CNM onto Celgard separators. Reproduced with permission from ref.
159 Copyright 2021 John Wiley and Sons. (b) Energy barrier of polysulfide migration obtained from DFT calculations. (c) The charge density difference of
Li
2
S
6
on the SAF-3 skeleton; purple colour represents the electron-rich zone and yellow colour represents the electron-deficient zone, respectively. (d)
Charge transfer analysis between Li
2
S
6
and SAF monomers and/or skeletons. Reproduced with permission from ref. 692 Copyright 2021 John Wiley and
Sons. (e) Schematic of the fabrication process to produce MOF@GO separators. The MOF nanoparticles and introduced GO laminates synergistically
comprise a MOF@GO separator. (f) An illustration of the microporous crystalline structures (HKUST-1). The homogeneous coordinated structures are
depicted as sticks, whereas the pores are highlighted in a space-filling representation. Reproduced with permission from ref. 517 Copyright 2016.
Springer Nature.
Chem Soc Rev Review Article
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2024
its redox conversion. As a result, SAF-3 modified batteries can
work well in a wide temperature range of 40 to 60 1C, showing
all-weather battery performance.
Moreover, MOFs with SNPCs can function as a sieve to
separate specific ions from an ionic solution based on their
sizes and shapes, making them akin to an ionic sieve. MOF-
based materials possessing a significant surface area and well-
structured pores with adjustable porosity are ideal choices as
ionic sieves for reducing the migration of polysulfide ions.
Thus, MOF modified separators are tailored for lithium–sulfur
batteries. This MOF acts as a battery separator by selectively
filtering Li
+
ions while preventing polysulfides from passing
through. Cu
3
(BTC)
2
(HKUST-1) is selected as the MOF to con-
struct the MOF@GO separator due to its three-dimensional
channel structure (Fig. 21e), which contains highly ordered
micropores. These micropores have a size window of about 9 Å
(Fig. 21f), making them significantly smaller than the dia-
meters of lithium polysulfide (Li
2
S
n
,4onr8). This property
makes it ideal for blocking polysulfides. The MOF@GO separa-
tor has demonstrated high efficiency in blocking polysulfides
and remarkable stability over long-term cycling in a lithium–
sulfur battery.
517
Advanced separators are beneficial for battery performance.
However, there are still many issues that separators cannot fully
solve. Due to the inevitable ion concentration gradient during
battery cycling, there can be uneven ion transport efficiency,
which can lead to dendrite formation on the anode side. In
addition, in metal–chalcogen battery systems, although multi-
functional separation membranes can block polysulfide spe-
cies, the unstable polysulfides species can still result in their
migration across the separator, which leads to irreversible loss
of active materials and continuous degradation of battery
performance. Therefore, embedding an interlayer is crucial
for metal-ion batteries with long cycle life, metal anode bat-
teries, and metal–chalcogen batteries. The Manthiram group
first proposed the important concept of ‘‘interlayer’’ in 2012
(Fig. 22a). Since then, extensive research has been conducted
on interlayer materials resulting in significant improvements.
Here, the interlayer of SNPCs plays an important role. In 2012,
Yu et al. first prepared the SNPCs hard carbon (HC) to encap-
sulate Li metal in these SNPCs (Fig. 22b and c), which can
prevent the direct contact between electrolyte and anode Li
while providing large accommodation space for Li storage,
restricting the dendrite growth of Li metal.
392
Recently, a
MoS
2
nanosheet has been used as an interlayer loaded on the
Celgard separator, seving as an effective polysulfide barrier,
resulting in high Coulombic efficient cycling stability
(Fig. 22d).
254
It is because the interspace between MoS
2
can
serve as an ionic sieve to selectively allow Li ion transport while
suppressing the shuttle effect. It was also found that chemical
reactions could occur between polysulfide and tungsten at the
edge of the micron-sized WS
2
wafer, effectively improving the
performance of Li–S cells.
704
When SNPC interlayers are
applied on the anode side, the interlayers are commonly
anchored to the separator facing the anode side or grown on
the surface of the anode. SNPCs can guarantee the smooth ion
transfer and avoid the corrosion reaction between the electro-
lyte and anode side. Moreover, preparing inherent regular
SNPCs on the SEI membrane could lead to uniform cation flux
at the electrode/electrolyte interface, causing uniform metal
deposition. A COF-based ultrafine artificial SEI was fabricated
by Guo’s group to regulate LiF formation and distribution in
the SNPCs in COFs. As illustrated in Fig. 22e, the high Li
+
affinity of COFs facilitates the adsorption of lithium salts from
the dilute electrolyte, resulting in an increase in the local
electrolyte concentration by an electrical double layer (EDL)
around the COF skeletons. Upon electrochemical lithiation of
the COF film during the first cycle, the reductive decomposition
of the strongly bound Li-salts locally generates uniformly dis-
tributed LiF domains. Such a homogeneous COF-LiF film
serves as a desirable SEI. The discrete LiF grains are confined
within the periodic microporous frameworks, endowing the
lithium/electrolyte interphase with considerably enhanced pas-
sivation and mechanical properties against dendrite growth
and electrolyte penetration. Meanwhile, the lithophilic moiety
of COFs facilitates Li
+
ion transport through the periodic COF
skeletons. Consequently, long-cycling stability of symmetric
lithium metal batteries has been attained in both ether- and
carbonate-based electrolytes.
705
Qiu’s research group prepared
a composite interlayer composed of carbon nanofibers (CNFs)
grown from a metal–organic framework (ZIF-67) enriched with
Co (Fig. 22f).
706
It is worth noting that the ZIF/CNF composite
interlayer can synergistically achieve physical blocking and
chemical capture capabilities, thereby inhibiting the dissolu-
tion of polysulfides and mitigating the shuttle effect. In addi-
tion, the three-dimensional fibrous network provides an
interconnected conductive framework between each ZIF
micro-reactor, facilitating rapid electron transfer during the
cycling process, which contributes to excellent rate and cycling
performance. The interlayer is also used in quasi-solid-state
electrolyte to reduce the loss of liquid electrolyte by adjusting
the solvation environment of moving ions in the layer. Thus,
Kim et al. introduce a novel approach to enhance the stability
and performance of Li metal batteries (LMBs). They present a
composite layer consisting of a single-ion-conducting ceramic
electrolyte (S-CE) and a single-ion-conducting polymer (S-PE)-
based gel electrolyte (S-GE).
518
This composite layer, referred to
as S-CE/S-GE, acts as a modulator for the reactivity of liquid
electrolyte components. By applying the S-CE/S-GE layer onto a
Li metal electrode, the loss of liquid electrolyte on the electrode
is significantly reduced. This layer also improves the cycling
stability of LMBs by altering the solvation environment of Li
+
ions within the layer (as shown in Fig. 22g). Interestingly, this
effect is not observed when using a composite layer comprising
S-CE and a bi-ion-conducting polymer (B-PE)-based gel electro-
lyte (B-GE). Furthermore, the researchers demonstrate that the
S-CE/S-GE layer has an impact on the morphology of Li anodes
and the stability of the cathode–electrolyte interface in Li metal
batteries. Leveraging the benefits of the S-CE/S-GE composite
layer, they successfully developed an energy-dense Li metal
pouch (Fig. 22h), which exhibits operation for over 400 cycles
with a low mass of electrolyte content of 2.15 g Ah
1
. The
Review Article Chem Soc Rev
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
This journal is © The Royal Society of Chemistry 2024 Chem. Soc. Rev.
careful engineering of the solvation structure and the resulting
outcomes provide valuable insights into the control of inter-
facial reactions and offer a practical methodology to enhance
the performance of LMBs.
4. Practical challenges
As comprehensively discussed above, electrochemical coupling
of SNPCs in a wide variety of electrode materials has been
crucial to enhance the performance of the batteries, which is
Fig. 22 (a) Schematic configuration of the cell with a bifunctional microporous carbon interlayer inserted between the sulphur cathode and the
separator. (b) SEM image of the surface of the MCP and (c) transmission electron microscope image of the microporous carbon particles. Reproduced
with permission from ref. 392 Copyright 2012 Springer nature. (d) Schematic cell configuration of Li–S batteries using the MoS
2
/Celgar separator.
Reproduced with permission from ref. 254 Copyright 2017 John Wiley and Sons. (e) Schematic illustration of the electrochemical preparation of the
COF-LiF hybrid interphase layer. Reproduced with permission from ref. 694 Copyright 2016. Springer Nature. (f) Mechanism of physical barrier and
chemical adsorption for polysulfides by the ZIF/CNFs interlayer. Reproduced with permission from ref. 695 Copyright 2021 Elsevier. (g) Schematic
illustration of the effects of the S-CE/B-GE and S-CE/S-GE composite layers on the cycling stability of liquid electrolyte-based LMBs and (h) cycling
stability with effect of S-CE/S-GE composite layers. Reproduced with permission from ref. 518 Copyright 2023 Springer Nature.
Chem Soc Rev Review Article
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2024
anticipated to be widely recognized and further explored in
future research studies. Despite significant advances in the past
few years, the employment of SNPCs in rechargeable batteries
is still in its infancy. To pave the way for future investigations, it
is essential to emphasize the practical challenges for SNPCs at
the current stage, as presented in Fig. 23.
4.1. Angstrom resolutions
The understanding of SNPCs has long been scarce, and one of
the major obstacles is the resolution limitations of traditional
characterization methods. To acquire accurate structural infor-
mation of the SNPCs, angstrom scale resolution is required.
The surface morphology of the samples is generally obtained
using a scanning electron microscope (SEM) which collects
scattered electrons for analysis, and the optimal spatial resolu-
tion achievable is about 0.5 nm, while transmission electron
microscopy (TEM), which collect transmitted electrons for
analysis, offers a much better spatial resolution with more
structural information such as lattice fringes and crystal struc-
tures. TEM can analyze the elemental composition of the
samples at high resolution when coupled with energy-
dispersive X-ray spectroscopy (EDS) or electron energy loss
spectroscopy (EELS); the former is particularly sensitive to
heavier elements while the later tends to work best for elements
with low atomic numbers. Whereas the specimen for TEM
analysis should be very thin (o150 nm) to permit electron
transmission, and even below 30 nm is generally required for
high resolution images. Scanning transmission electron micro-
scopy (STEM) can be deemed as the combination of SEM
and TEM, which can analyze both scattered and transmitted
electrons; it requires stable room environments (limited
temperature fluctuations, electromagnetic waves, acoustic
waves, etc.) and can deliver an optimal spatial resolution of
0.5 Å.
707
With the advancement of detectors and computational
power, 4D STEM with beam rasters across a 2 dimensional
region in real space has been invented.
708
It does not necessa-
rily result in higher resolution for 4D STEM; however it can be
extremely helpful for specimens with distorted structures due
to high strain. It is worth noting that, it generally requires to
apply a high dose of electron on a small sampling area to get
enough yield of electrons for high resolution analysis, whereas
many samples are less tolerable to electrons and can get
damaged by the focused electrons. Cryogenic electron micro-
scopy (cryo-EM) is another extension of TEM technology, which
can operate at cryogenic temperatures to minimize the sample
movement or damage. Cryo-EM has been widely used for
structure determination of biomolecules, which has recently
been utilized for battery materials, offering significant contri-
butions to the field.
709
The invention of cryo-EM enabled the
acquisition of atomic structures for a wide variety of sensitive
battery materials at the angstrom scale, such as, air and
electron sensitive Li/Na metal, air and electron sensitive SEI
layers, and volatile S elements.
710
In addition to various electron microscopy techniques, other
morphology and structural information from diverse character-
ization techniques are also essentially required. X-ray diffrac-
tion (XRD) has been a powerful tool to characterize the lattice
patterns of nanocrystals at the angstrom scale, while the lattice
distortion may cause broadening and overlap of the diffraction
peaks.
711
High spatial and angular resolution XRD can be
achieved by reducing the beam divergence in high photon
fluxes synchrotron sources. For instance, synchrotron XRD with
Fig. 23 The practical challenges associated with SNPCs for rechargeable batteries.
Review Article Chem Soc Rev
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
This journal is © The Royal Society of Chemistry 2024 Chem. Soc. Rev.
a high resolution of 1.9 Å has been achieved, which could
identify light elements such as oxygen and water molecules.
712
Besides, X-rays are suited for superficial analysis of shallow
depths or thin specimens, while neutron diffraction with its
high penetration depth is suited for bulk samples.
713
Other
angstrom characterization techniques, such as, atomic force
microscopy (AFM) and tip-enhanced Raman spectroscopy
(TERS) should be further investigated for SNPCs.
714
Since the
SNPCs cover a wide range of material families with diverse
physicochemical properties, it is necessary to choose suitable
specimens for characterization of angstrom resolutions to
acquire the desired information.
4.2. Operando characterization
For the electrochemical coupling of SNPCs in various batteries,
dynamic evolution of various species during the electrochemi-
cal reactions is of particular interest. Whereas most of the
results were achieved through post-mortem analysis, for which
the composition of the electrodes may have changed during the
sampling process. Therefore, it is essential to develop various
operando or in situ characterization techniques to investigate
the dynamic electrochemical evolutions under real time work-
ing conditions. Here, the high-resolution in situ transmission
electron microscope (TEM) electrochemical experimental plat-
form is very important for observing battery reactions. This
technology allows direct visualization and timely discovery of
changes in materials during the charging and discharging
process, helping researchers to deeply understand the mecha-
nism of SNPCs for ion storage. In addition, high-resolution
in situ TEM can also observe the phase change process of the
cathode material during battery operation, which helps
researchers understand the working mechanism of the battery
and make timely adjustments. The development of high-
resolution in situ TEM allows researchers to observe the struc-
tural changes of materials during charging and discharging at
the atomic level in real time. This is crucial for understanding
the mechanism of battery reactions and optimizing battery
adjustments. Compared with traditional post-reaction charac-
terization, high-resolution in situ TEM can precisely control and
capture the dynamic changes of materials, revealing the rela-
tionship between the structure and performance. However, the
current in situ TEM technology still faces great challenges; first,
it is difficult to carry out high-rate (atomic level) characteriza-
tion; second, the ability to combine other technologies such as
EDS and EELS is still weak. The high-resolution in situ TEM
would become a powerful tool, not only capturing structural
information of nanomaterials in specific environments, but
also guiding the design of functional materials. In addition,
another feasible option is to use a high-throughput synchrotron
light source to penetrate the seal of the battery and generate
enough signals for analysis, for instance, synchrotron-based
XRD, synchrotron-based X-ray absorption spectroscopy
(XAS),
715
and synchrotron-based scanning transmission X-ray
microscopy (STXM).
716
In situ synchrotron XRD has been widely
used for various battery systems to investigate the lattice
parameters, phase transitions, and strain evolution of electrode
materials under real time operando conditions, which are
extremely useful for an in depth understanding of the working
mechanisms. In situ synchrotron XAS could provide the oxida-
tion state and coordination environment of the elements (X-ray
absorption near edge structure, XANES), and also the coordina-
tion numbers, bonding distances, and the chemical identity of
nearest neighbours (X-ray absorption fine structure, EXAFS).
717
In situ synchrotron STXM could achieve site-specific informa-
tion on chemical and structural changes of the electrode
materials within single particles at high spatial resolution
(B20 nm), therefore, it is powerful for understanding the
working mechanisms of the electrode materials under operando
conditions. Other synchrotron-based characterization techni-
ques, for instance, synchrotron micro-computed tomography
(MCT) while unable to achieve high resolution at the angstrom
scale, could also be further explored for SNPCs due to their
capability of acquiring various structural information under
operando conditions.
4.3. Atomic perceptions
Up to now, most of the reports on SNPCs for battery applica-
tions are focused on the microstructures and electrochemical
performance of the electrodes, whereas an in depth under-
standing of the working mechanisms at the atomic level is
insufficient. Accompanied by the abovementioned angstrom
resolution and operando characterization studies, it is much
more desirable to decipher the working mechanism with
atomic perceptions. For instance, the encapsulation of S in
SNPCs could alter the atomic configuration from cyclic S
8
to
short chain S
2–4
species, which could bypass the generation of
soluble long chain polysulfides to minimize S dissolution. This
is a good example of the process from quantitative change to
qualitative change, in which the impact of the SNPCs should be
carefully considered with atomic perceptions. Besides, the
atomic configurations of the doping elements in SNPCs, e.g.,
S doping at different sites in hard carbon, could result in
significantly different electrochemical performance due to dif-
ferent working mechanisms, further verifying the significance
of angstrom resolution characterization studies and atomic
perceptions. For the electrochemical coupling of SNPCs in
various batteries, it is highly important to have a closer look
at the working mechanisms from an atomic perspective. Espe-
cially, the solvation structure of the ions, the composition of
the SEI layers, the influence of the doping elements with
different atomic configurations, the atomic configuration of
the species restricted in the SNPCs, and the interaction of
various species with the building blocks of SNPCs should be
investigated with atomic perceptions to explore the underlying
working mechanisms.
4.4. Theoretical calculations
With the significant advancement of theoretical calculations in
the last two decades, many key properties of the batteries can
be accurately predicted. The capabilities of theoretical calcula-
tions for batteries including theoretical capacity, equilibrium
voltage, ion diffusion pathways, electronic conductivity, rate
Chem Soc Rev Review Article
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2024
capability, thermal stability, etc.
718
These properties are invalu-
able for battery research community, offering an alternative
‘‘prediction and validation’’ route instead of ‘‘trial and error’’
routine for experimental research. Therefore, the potential of
theoretical calculations for electrochemical coupling of SNPCs
in carious batteries should be further explored to enhance the
fundamental understanding of the working mechanisms. For
instance, selective ion transport in SNPCs of MOF based SSEs
has been validated through density functional theory (DFT)
calculations and molecular dynamics (MD) simulations, which
are essential for the atomic understanding of the working
mechanism.
719
While theoretical calculations for various bat-
tery materials have been frequently carried out for crystalized
inorganic materials, modelling and calculation of organic and/
or amorphous materials have been challenging. The SEI layers
on the anode surfaces which are generally composed of organic
and inorganic composites are very meaningful yet challenging
for DFT calculations. As for the SNPCs, one must consider the
size effect of the nano particles, for which the properties may be
dramatically changed at the subnanometer scale. For example,
nanosized LiFePO
4
could deliver ultrafast Li
+
extraction, for
which the low-energy surfaces of the nanocrystals account for
up to 85% of the total surface area.
720
Besides, calculations
for solid-state materials with known crystal structures have
been widely achieved, while those with multiple phases and/or
physical fields are difficult, e.g., the interfaces of the electrodes
with complex interactions such as electric field, electrostatic
field, steric hindrance, solvation structure, ion dissociation,
electrochemical voltage window, etc.
Furthermore, with the development of computational power
and artificial intelligence, a new branch of computation known
as ‘‘machine learning’’ has recently surged. Machine learning is
a protocol to feed the program with plenty of known results for
the computer and establish an approximate model between
input and output, with which artificial intelligence could learn
and discover an optimal algorithm to summarize the results
with minimum deviation.
721
After numerous optimizations, the
algorithms can be used to predict the results for unknown
species or systems. It is remarkable that we can find the
algorithms before understanding the underlying correlations
between the properties and the performance, the algorithms in
turn could be helpful to discover the decisive parameters for
the systems. The capabilities of machine learning should be
further explored for SNPCs in batteries, which could certainly
lead to a new era for close combination of theories with
experiments by data-driven computations.
5. Conclusion and perspectives
To meet the growing demand for batteries, the development of
new materials that improve the efficiency of energy conversion
and storage systems is crucial. The materials with SNPCs offer
opportunities in energy conversion and battery storage applica-
tions due to their unique physio-chemical properties. As illu-
strated in Fig. 24, the most straightforward advantage of the
SNPCs is that they could achieve regulated ion transportation,
facilitate selective and fast ionic transport through the size
effect or electrostatic effect through the SNPCs.
435,722–724
Besides, the SNPCs with carbon layers or conductive polymers
can serve as an electrically conductive matrix, realizing superior
electrical conductivity. The combination of fast ion transfer and
high overall electrical conductivity could contribute to high-rate
capabilities.
392,725
In addition, some SNPCs as artificial SEIs or
protecting layers could relieve the volume change of the elec-
trode materials as a result of the suppression effect, to
enhance the cycling stability of the electrodes.
726,727
The SNPC
protecting layers can also deliver good ion and electron
conductivity, to enhance the rate performance of the
electrodes.
728,729
Furthermore, by introducing SNPCs into the
materials, higher surfaces area with additional active sites can
be achieved; therefore the reversible specific capacity can be
increased.
730,731
Noticeably, the void formation should be
avoided during the SNPCs fabrication process. The generation
of voids will greatly affect the inherent electrical conductivity of
the matrix. When the void exists, the electrons transformation
will be greatly suppressed. The electrons must achieve more
energy to overcome the transfer barrier created by the voids. In
a word, the SNPCs in the battery materials can enhance the rate
capability, enhance the cycling stability, and increase the
reversible specific capacity of the anodes. Thereafter, the sig-
nificance of SNPCs for various anode materials should be
further explored in the following research.
This review comprehensively covers all aspects of SNPCs and
their involvement in electrochemical reactions within batteries
for the first time. We first define SNPCs and then, we rationally
sorted the families of SNPCs materials and summarize the
unique physicochemical properties of SNPCs. With respect to
the electrochemical applications of SNPCs in rechargeable
batteries, we have systematically summarized the electroche-
mical reactions occurring on various components of the battery
in SNPCs. The challenges for the practical application of SNPCs
in batteries have also been discussed in detail. Furthermore, to
further explore the huge potentials of SNPCs for rechargeable
batteries, we have summarized the perspectives for future
research as followed.
Synthesis of SNPCs
With the highly sought-after attributes of simplicity, environ-
mental friendliness, and controllability of SNPCs structure
synthesis materials, an increase in supply will inevitably lead
to a wider range of applications. Compared to the synthesis of
materials with larger pores, mesopores, and micropores, the
synthesis of SNPCs structures requires carefully designed pre-
paration schemes because their sensitivity to reaction condi-
tions depends on each composition and intermediate phase. In
most methods, such as template methods, they may not be
applicable to the reconstruction of controllable SNPCs. In
addition, the size, structure, and interaction of the components
will greatly affect the synthesis of SNPCs structures. Consider-
ing these factors, it is crucial to find a template-free, simple,
and effective technique to control the synthesis of SNPC
Review Article Chem Soc Rev
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
This journal is © The Royal Society of Chemistry 2024 Chem. Soc. Rev.
materials, like the use of metal–organic frameworks in chem-
istry. In energy conversion and storage devices, controlling the
overall particle size, surface structure, morphology, packing
density, orderliness, and high density is also crucial. At the
same time, favorable structural parameters are different for
different applications. Methods for systematically controlling
and optimizing these parameters for specific application sys-
tems have not yet been developed.
High specific surface areas and porosity are ideal conditions
for improving activity. Nevertheless, these advantages provide
more opportunities for adverse reactions. This is particularly
evident in batteries, where the combination of a high specific
surface area increases reactivity with the electrolyte and leads to
uncontrolled solid electrolyte interface reactions. Additionally,
in relation to volumetric energy and power density (reflecting
how much and how fast energy can be stored in a unit volume
of packaging equipment), the low packing density of SNPC
materials may be another disadvantage. Furthermore, SNPC
materials with nanoscale frameworks and high surface energy
often have lower thermal stability, which affects the catalytic
performance at high temperatures. Thus, SNPC materials with
ultra-high specific surface area will further increase the burden
of electrolyte infiltration. SNPCs materials with ultra-high
specific surface area typically have lower density. Therefore,
the key future research direction is how to use less electrolyte to
wet high-load mass electrodes to achieve high utilization of
active materials. In addition, the mechanical stability of the
materials is also crucial, as excessive porosity may pose a risk of
local collapse, and more inconsistent porosity may severely
affect the improvement in battery performance. In summary,
the synthesis of future sub-nanometer pores should be devel-
oped based on applications, and the development of sub-
nanometer pores should be beneficial for the reactions
involved in those applications. For batteries, the future SNCP
designs should prioritize characteristics such as ease of synthe-
size, high-density, controllable, high toughness and mechan-
ical performance, easy wetting by electrolyte, better
compatibility with electrolyte or chemical materials, non-toxic
and environmentally friendly, and low-cost, similar to current
nanometer pore designs.
Stability of materials with SNPCs
Due to the large surface area of SNPCs, SNPCs can theoretically
provide more space to accommodate active materials or
cations, which makes the matrix materials more resilient to
volume changes and preserve structural stability. However, it is
intractable to maintain the SNPC stability for intercalation-type
electrodes due to intrinsic working mechanisms. Taking
intercalation-type electrodes as examples, the structure will
undergo transformation due to intercalation/deintercalation
during charge/discharge processes, which make the lattice ions
deviate from their original positions, leading to potential risks
of structural collapse.
732,733
In terms of conversion-type electro-
des, especially for metal–sulfur batteries, the structural stability
of the SNPCs is relatively hard to maintain. As a S host, the
SNPCs have to sustain large volume deformation due to the
severe volume expansion of S during redox reactions, causing
serious structure damage or even physical disintegration.
734
In
addition, whether it is an intercalation-or alloy-type anode,
unevenly sized and inappropriate SNPCs would result in non-
uniform deposition of metal ions during the charging process.
As the cycles progress, the structure of the anode electrode will
be severely damaged.
735,736
Overall, the structural stability of
SNPCs is wildly considered one of the main research directions.
Fig. 24 Summary of the versatile functions of the SNPCs for rechargeable batteries.
Chem Soc Rev Review Article
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2024
Therefore, more research attention should be given to enhance
stability of materials with SNPCs.
Characterization of SNPCs
This includes the characterization of the sub-nanometer pores
themselves as well as the characterization of electrochemical
reactions in sub-nanometer pores. Current technological lim-
itations result in material characterization only reaching the
nanometer level. Therefore, upgrading and developing new
generation characterization devices is crucial and urgent.
Three-dimensional atom probe tomography (APT) can provide
visual and quantitative analysis of materials. Visual analysis
includes observing the internal grain boundaries, phase bound-
aries, structural interfaces, as well as dislocations and defects at
different locations. Quantitative analysis includes determining
the composition of materials, calculating the quantity, density,
and volume of precipitation and segregation at interfaces or
grain boundaries. Moreover, nanoscale electrochemical reac-
tions need to be characterized, such as the confinement of
SNPCs, ion diffusion, sub-nanometer solvent effects, etc. How
to characterize these effects is a top priority for future research.
For this, more advanced characterization techniques need to be
combined with advanced electron microscopy for further judg-
ment. For instance, wide-angle X-ray scattering (WAXS) mea-
surements provide a spatial resolution in the range of 0.2 nm–
1 nm; By combination of WAXS and SXAS, the capability can be
extended to determine the microstructural features for porous
materials with SNPCs.
In addition to the characterization of the nanopores them-
selves, further development is needed for the characterization
of electrochemical reactions within nanopores. Due to techni-
cal limitations, it is difficult to directly obtain external char-
acterization of the electrochemical reactions of SNPCs, such as
transmission electron microscopy (TEM), X-ray photoelectron
spectroscopy (XPS), scanning electron microscopy (SEM), XRD,
Raman spectroscopy, Fourier transform infrared spectroscopy
(FTIR), etc. To fully understand the integrated multi-scale
regulation mechanism, several challenges related to character-
ization techniques need to be addressed. The first challenge is
the time scale and operating environment. Although offline
characterization studies provide valuable information at rela-
tively low cost, their limitations restrict their application: on
one hand, there is a lack of direct observation of the movement
processes of molecules, atoms, ions, etc., as well as the direct
reaction intensity. On the other hand, due to the sensitivity of
electrolytes and metal ions to air and moisture, offline char-
acterization cannot accurately represent the reaction processes
in batteries. Therefore, in situ/real-time characterization tech-
niques are crucial for a deep understanding of the solvation
chemistry in various rechargeable batteries. In addition to
using more advanced high-resolution electron microscopy,
more instrumental methodologies need to be developed, such
as how to further transfer the high-resolution electron micro-
scopy imaging technology to digital intensity, which can help
us improve this aspect. Through this means, not only can we
observe whether local reactions are intense or not, but we can
also observe the movement and reactions of ions, atoms, and
molecules. Combined with in situ techniques, this will elevate
our understanding of battery reactions from the nano scale to
the sub-nano scale.
Advanced theoretical calculation and simulation techniques
Here, we propose the use of material knowledge supported
machine learning (MIML) to enhance current material compu-
tation methods.
737
MIML combines machine learning techni-
ques with explicit prior material knowledge. In fact, many
successful cases in materials science already utilize specific
forms of MIML, such as deep density functions, chemical
synthesis planning, and the use of deep neural networks for
ab initio calculations in multi-electron systems. By employing
MIML-based algorithms, the computational workflow becomes
partially interpretable, and the need for extensive training data
is greatly reduced due to the incorporation of prior knowledge
or mathematical models. In some cases, it may even achieve
zero-data training. Our goal, from the perspective of multiscale
simulation in material design, is to develop a general partial
differential equation (PDE) solver for subnanometer-scale
simulations based on MIML. Additionally, we aim to propose
a neural network alternative model for solving ab initio Schro
¨-
dinger equations with built-in physical constraints in the
future. Overall, MIML-based material computation platforms
have the potential to extend multiscale simulations to larger
systems with higher accuracy.
Furthermore, this is a comprehensive platform based on
‘‘artificial intelligence + big data’’. Due to its efficiency, artificial
intelligence (AI) (whose core is machine learning (ML), high-
throughput simulation, and experimentation) and big data
technology have attracted more and more attention in molecu-
lar design, material development, performance prediction,
mechanism analysis, relationship mining, and other fields.
Therefore, we propose a comprehensive platform based on
‘‘artificial intelligence + big data’’. According to the needs,
high-throughput computation and simulation based on
machine learning are first used to screen and simulate feasible
molecules and ions with suitable pore sizes, and then high-
throughput synthesis, characterization, and battery testing are
carried out by mobile robot technicians. All the results obtained
in the above process are stored in the database, providing
enough data for machine learning to optimize the design and
performance prediction of SNPCs or sub-nanometer structures.
Finally, high-performance ideal materials are obtained effi-
ciently and quickly. In addition, this comprehensive platform
is an efficient tool for understanding sub-nanometer scale
mechanisms or reactions, such as electrochemical reactions
in SNPCs that are still difficult due to insufficient experimental
techniques.
To better regulate and improve batteries fundamentally
and enhance battery performance, a deeper understanding of
the structure, electron, surface characteristics, and transfer
mechanisms of SNPCs materials at the level of battery and
electrochemistry need to be achieved. Given that energy con-
version and storage devices will be at the forefront in the
Review Article Chem Soc Rev
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
This journal is © The Royal Society of Chemistry 2024 Chem. Soc. Rev.
coming years, SNPCs materials are bound to become the next
material trend. Of course, this not only depends on the devel-
opment of materials but also relies on the innovation and
manufacturing of advanced characterization equipment. There-
fore, similar to nanomaterials, it is necessary to conduct
fundamental research, manufacturing, and techno-economic
analysis on SNPCs materials in order to successfully produce and
characterize SNPC materials in a controllable manner and inte-
grate them with cost-effective battery industrial systems. Although
there seem to be many difficulties, we still believe that through
collaboration and interdisciplinary efforts, there will be more and
more exciting results, enabling sub-nanotechnology to be widely
applied like current nanotechnology.
Abbreviations
SNPCs Subnanometer pore/channels
MOFs Metal–organic frameworks
COFs Covalent organic frameworks
POCs Porous organic cages
TMDs Transition metal dichalcogenides
TMOs Transition metal oxides
CNTs Carbon nanotubes
GAs Graphene analogues
g-C
3
N
4
Graphitic carbon nitrides
MWCNTs Multi-walled carbon nanotubes
SWCNTs Single-walled carbon nanotues
DWCNTs Double-walled carbon nanotubes
TMPs Transition metal phosphates
MXenes Transition metal nitrides/carbides
TMDs Transition metal dichalcogenides
POCs Porous organic cages
SSIC sodium super ionic conductor
PBAs Prussian blue analogues
POCs Porous organic cages
LIBs Lithium-ion batteries
SIBs Sodium-ion batteries
PIBs Potassium-ion batteries
LSBs Lithium–sulfur batteries
SSBs Sodium–sulfur batteries
PSBs Potassium–sulfur batteries
SEI Solid–electrolyte interphase
CMPs Conjugated microporous polymers
QSSEs Quasi-solid-state electrolytes
GPEs Gel polymer electrolytes
PIMs Polymer of inherent microporosity
PAFs Porous aromatic frameworks
POFs Porous organic frameworks
SPEs Solid polymer electrolytes
HR-TEM High resolution transmission electron microscopy
HAADF-STEM High angle annular dark field scanning trans-
mission electron microscopy
ABF-STEM Annular bright field scanning transmission
electron microscopy
iDPC Integrated differential phase contrast
XPS X-ray photoelectron spectroscopy
HAXPES Hard X-ray photoelectron spectroscopy
XANES X-ray absorption near edge structure
EELS Electron energy loss spectroscopy
XRPD X-ray powder diffraction
NPD Neutron powder diffraction
EDT Electron diffraction tomography
SAED Selected area electron diffraction
PDF Pair distribution function analysis
RIXS Resonant inelastic X-ray scattering
NMR Nuclear magnetic resonance spectroscopy
WAXS Wide angle X-ray/neutron scattering
XCT X-ray computed tomography
XRD-CT X-ray diffraction computed tomography
APT Atom probe tomography
NMRC Nuclear magnetic resonance nano pore analysis
Author contributions
Conceptualization, Y. -J. Lei., W. -H. Lai., Y.-X. Wang. and
G. Wang.; visualization, Y. -J. Lei., L. Zhao., Z. Huang, and
P. Jaumaux.; writing original draft, Y.-J. Lei., L. Zhao., Z.
Huang., and P. Jaumaux.; writing review and editing, Y.-J. Lei.,
L. Zhao., W. -H. Lai., B. Sun., Y.-X. Wang., and G. Wang.;
funding acquisition, W. -H. Lai., Y. -X. Wang., B. Sun., and G.
Wang.; and supervision, G. Wang.
Conflicts of interest
There are no conflicts to declare.
Acknowledgements
We would like to acknowledge the support from the Australian
Research Council (ARC) through the Discovery Projects
(DP210101389 and DP220103301), the ARC Future Fellowship
(FT180100705), the ARC Discovery Early Career Researcher Award
(DE220101113), the ARC Linkage project (LP200200926), and the
ARC Research Hub for Integrated Energy Storage Solutions
(IH180100020). The authors thank Dr Tania Silver for critical
discussion.
References
1 Z. Zhang, X. Yang, P. Li, Y. Wang, X. Zhao, J. Safaei,
H. Tian, D. Zhou, B. Li, F. Kang and G. Wang, Adv. Mater.,
2022, 34, 2206970.
2 F. Liu and Z. Fan, Chem. Soc. Rev., 2023, 52, 1723–1772.
3 D. Luo, M. Li, Q. Ma, G. Wen, H. Dou, B. Ren, Y. Liu,
X. Wang, L. Shui and Z. Chen, Chem. Soc. Rev., 2022, 51,
2917–2938.
4 Y. Yang, P. Li, X. Zheng, W. Sun, S. X. Dou, T. Ma and
H. Pan, Chem. Soc. Rev., 2022, 51, 9620–9693.
5 J. M. Tarascon and M. Armand, Nature, 2001, 414, 359–367.
Chem Soc Rev Review Article
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2024
6 Y. Zhao, J. Zhang, X. Guo, X. Cao, S. Wang, H. Liu and
G. Wang, Chem. Soc. Rev., 2023, 52, 3215–3264.
7 Y. Wan and D. Zhao, Chem. Rev., 2007, 107, 2821–2860.
8 X. Li, F. Zhang and D. Zhao, Chem. Soc. Rev., 2015, 44,
1346–1378.
9 T. Zhao, A. Elzatahry, X. Li and D. Zhao, Nat. Rev. Mater.,
2019, 4, 775–791.
10 M. R. Palacı
´n, Chem. Soc. Rev., 2009, 38, 2565–2575.
11 W. Song, Q. Jiang, X. Xie, A. Brookfield, E. J. L. McInnes,
P. R. Shearing, D. J. L. Brett, F. Xie and D. J. Riley, Energy
Storage Mater., 2019, 22, 441–449.
12 S. Li, J. Huang, Y. Cui, S. Liu, Z. Chen, W. Huang, C. Li,
R. Liu, R. Fu and D. Wu, Nat. Nanotechnol., 2022, 17,
613–621.
13 L. Kong, M. Zhong, W. Shuang, Y. Xu and X.-H. Bu, Chem.
Soc. Rev., 2020, 49, 2378–2407.
14 C. Liu, F. Li, L.-P. Ma and H.-M. Cheng, Adv. Mater., 2010,
22, E28–E62.
15 Y. Zhai, Y. Dou, D. Zhao, P. F. Fulvio, R. T. Mayes and
S. Dai, Adv. Mater., 2011, 23, 4828–4850.
16 H. Wang, F. Sun, Z. Qu, K. Wang, L. Wang, X. Pi, J. Gao and
G. Zhao, ACS Sustain. Chem. Eng., 2019, 7, 18554–18565.
17 C. Chen, Y. Huang, Y. Zhu, Z. Zhang, Z. Guang, Z. Meng
and P. Liu, ACS Sustain. Chem. Eng., 2020, 8, 1497–1506.
18 G. Cai, Y. Yin, D. Xia, A. A. Chen, J. Holoubek, J. Scharf,
Y. Yang, K. H. Koh, M. Li and D. M. Davies, Nat. Commun.,
2021, 12, 3395.
19 P. Wang, N. Kateris, B. Li, Y. Zhang, J. Luo, C. Wang,
Y. Zhang, A. S. Jayaraman, X. Hu, H. Wang and W. Li, J. Am.
Chem. Soc., 2023, 145, 18865–18876.
20 C. Wu, Y. Lei, L. Simonelli, D. Tonti, A. Black, X. Lu,
W.-H. Lai, X. Cai, Y.-X. Wang, Q. Gu, S.-L. Chou, H.-K.
Liu, G. Wang and S.-X. Dou, Adv. Mater., 2022, 34,
2108363.
21 M. Shen and H. Ma, Coord. Chem. Rev., 2022, 470,
214715.
22 E. Detsi, X. Petrissans, Y. Yan, J. B. Cook, Z. Deng,
Y.-L. Liang, B. Dunn and S. H. Tolbert, Phys. Rev. Mater.,
2018, 2, 055404.
23 A. D. Roberts, X. Li and H. Zhang, Chem. Soc. Rev., 2014, 43,
4341–4356.
24 V. Augustyn, J. Mater. Res., 2017, 32, 2–15.
25 Y. Guo, Y. Wei, H. Li and T. Zhai, Small, 2017, 13,
1701649.
26 Z. Liu, L. Qin, X. Cao, J. Zhou, A. Pan, G. Fang, S. Wang and
S. Liang, Prog. Mater. Sci., 2022, 125, 100911.
27 A. Van der Ven, J. Bhattacharya and A. A. Belak, Acc. Chem.
Res., 2013, 46, 1216–1225.
28 D. He, L. Zhang, T. Liu, R. Clowes, M. A. Little, M. Liu,
M. Hirscher and A. I. Cooper, Angew. Chem., Int. Ed., 2022,
61, e202202450.
29 B. Ni, Y. Shi and X. Wang, Adv. Mater., 2018, 30, 1802031.
30 P. Zhang, F. Wang, M. Yu, X. Zhuang and X. Feng, Chem.
Soc. Rev., 2018, 47, 7426–7451.
31 Y. Chen, C. Tan, H. Zhang and L. Wang, Chem. Soc. Rev.,
2015, 44, 2681–2701.
32 N. Liu, K. Xu, Y. Lei, Y. Xi, Y. Liu, N. Wang, Y.-X. Wang,
X. Xu, W. Hao, S. X. Dou and Y. Du, Small Struct., 2021,
2, 2100041.
33 X. Xiao, H. Wang, P. Urbankowski and Y. Gogotsi, Chem.
Soc. Rev., 2018, 47, 8744–8765.
34 C. Li, L. Liu, J. Kang, Y. Xiao, Y. Feng, F.-F. Cao and
H. Zhang, Energy Storage Mater., 2020, 31, 115–134.
35 D. S. Bethune, C. H. Kiang, M. S. de Vries, G. Gorman,
R. Savoy, J. Vazquez and R. Beyers, Nature, 1993, 363,
605–607.
36 M. T. Pettes and L. Shi, Adv. Funct. Mater., 2009, 19,
3918–3925.
37 J. P. Paraknowitsch and A. Thomas, Energy Environ. Sci.,
2013, 6, 2839–2855.
38 J. C. Charlier, Acc. Chem. Res., 2002, 35, 1063–1069.
39 A. Saha, B. J. Gifford, X. He, G. Ao, M. Zheng, H. Kataura,
H. Htoon, S. Kilina, S. Tretiak and S. K. Doorn, Nat. Chem.,
2018, 10, 1089–1095.
40 Y. Shim and H. J. Kim, ACS Nano, 2009, 3, 1693–1702.
41 X. Yi, X. Liu, R. Dou, Z. Wen and W. Zhou, J. Phys. Chem. C,
2021, 125, 22570–22580.
42 H. Tabassum, A. Mahmood, B. Zhu, Z. Liang, R. Zhong,
S. Guo and R. Zou, Energy Environ. Sci., 2019, 12, 2924–2956.
43 C. Fu, M. B. Oviedo, Y. Zhu, A. von Wald Cresce, K. Xu,
G. Li, M. E. Itkis, R. C. Haddon, M. Chi, Y. Han,
B. M. Wong and J. Guo, ACS Nano, 2018, 12, 9775–9784.
44 Y. Cao, M. Wang, H. Wang, C. Han, F. Pan and J. Sun, Adv.
Energy Mater., 2022, 12, 2200057.
45 H.-S. Xu, S.-Y. Ding, W.-K. An, H. Wu and W. Wang, J. Am.
Chem. Soc., 2016, 138, 11489–11492.
46 N. Huang, L. Zhai, D. E. Coupry, M. A. Addicoat,
K. Okushita, K. Nishimura, T. Heine and D. Jiang, Nat.
Commun., 2016, 7, 12325.
47 F. Yu, W. Liu, B. Li, D. Tian, J.-L. Zuo and Q. Zhang, Angew.
Chem., Int. Ed., 2019, 58, 16101–16104.
48 Y.Zeng,R.ZouandY.Zhao,Adv. Mater., 2016, 28, 2855–2873.
49 Z. Wang, J. Hu and Z. Lu, Batter. Supercaps, 2023,
6, e202200545.
50 L. Liu and A. Corma, Nat. Rev. Mater., 2021, 6, 244–263.
51 Y. Yang, X. Chu, H.-Y. Zhang, R. Zhang, Y.-H. Liu,
F.-M.Zhang,M.Lu,Z.-D.YangandY.-Q.Lan,Nat. Commun.,
2023, 14, 593.
52 W. Xin, J. Fu, Y. Qian, L. Fu, X.-Y. Kong, T. Ben, L. Jiang
and L. Wen, Nat. Commun., 2022, 13, 1701.
53 F. Xu, S. Jin, H. Zhong, D. Wu, X. Yang, X. Chen, H. Wei,
R. Fu and D. Jiang, Sci. Rep., 2015, 5, 8225.
54 C. Zhang, C. Lu, F. Zhang, F. Qiu, X. Zhuang and X. Feng,
J. Energy Chem., 2018, 27, 86–98.
55 X. Ma and T. F. Scott, Commun. Chem., 2018, 1, 98.
56 X. Zhan, Z. Chen and Q. Zhang, J. Mater. Chem. A, 2017, 5,
14463–14479.
57 R.-B. Lin, Z. Zhang and B. Chen, Acc. Chem. Res., 2021, 54,
3362–3376.
58 J. Wang, Y. Zhang, Y. Su, X. Liu, P. Zhang, R.-B. Lin,
S. Chen, Q. Deng, Z. Zeng, S. Deng and B. Chen, Nat.
Commun., 2022, 13, 200.
Review Article Chem Soc Rev
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
This journal is © The Royal Society of Chemistry 2024 Chem. Soc. Rev.
59 P. Yan, F. Ai, C. Cao and Z. Luo, J. Mater. Sci.: Mater.
Electron., 2019, 30, 14120–14129.
60 J. Lu, H. Zhang, J. Hou, X. Li, X. Hu, Y. Hu, C. D. Easton,
Q. Li, C. Sun, A. W. Thornton, M. R. Hill, X. Zhang,
G. Jiang, J. Z. Liu, A. J. Hill, B. D. Freeman, L. Jiang and
H. Wang, Nat. Mater., 2020, 19, 767–774.
61 P. Xiong, F. Zhang, X. Zhang, Y. Liu, Y. Wu, S. Wang,
J. Safaei, B. Sun, R. Ma, Z. Liu, Y. Bando, T. Sasaki,
X. Wang, J. Zhu and G. Wang, Nat. Commun., 2021,
12, 4184.
62 W. Li, R. Fang, Y. Xia, W. Zhang, X. Wang, X. Xia and J. Tu,
Batter. Supercaps, 2019, 2, 9–36.
63 S. Dang, Q.-L. Zhu and Q. Xu, Nat. Rev. Mater., 2017,
3, 17075.
64 P. Liu, J. Yan, Z. Guang, Y. Huang, X. Li and W. Huang,
J. Power Sources, 2019, 424, 108–130.
65 C. Gao, F. Lyu and Y. Yin, Chem. Rev., 2021, 121, 834–881.
66 W. Liu, P. Oh, X. Liu, M.-J. Lee, W. Cho, S. Chae, Y. Kim
and J. Cho, Angew. Chem., Int. Ed., 2015, 54, 4440–4457.
67 P.-F. Wang, Y. You, Y.-X. Yin and Y.-G. Guo, Adv. Energy
Mater., 2018, 8, 1701912.
68 K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang,
Y. Zhang, S. V. Dubonos, I. V. Grigorieva and A. A. Firsov,
Science, 2004, 306, 666–669.
69 Z. Xu, Z. Jiang, C. Kuai, R. Xu, C. Qin, Y. Zhang,
M. M. Rahman, C. Wei, D. Nordlund, C.-J. Sun, X. Xiao,
X.-W. Du, K. Zhao, P. Yan, Y. Liu and F. Lin, Nat. Commun.,
2020, 11, 83.
70 K. Ji, J. Han, A. Hirata, T. Fujita, Y. Shen, S. Ning, P. Liu,
H. Kashani, Y. Tian, Y. Ito, J.-I. Fujita and Y. Oyama, Nat.
Commun, 2019, 10, 275.
71 J. Xiao, M. Long, X. Li, H. Xu, H. Huang and Y. Gao, Sci.
Rep., 2014, 4, 4327.
72 Z. Hu, Q. Liu, S.-L. Chou and S.-X. Dou, Adv. Mater., 2017,
29, 1700606.
73 J. Mei, T. Liao, L. Kou and Z. Sun, Adv. Mater., 2017,
29, 1700176.
74 S. Wang, S. Zhao, X. Guo and G. Wang, Adv. Energy Mater.,
2022, 12, 2100864.
75 C. Eames and M. S. Islam, J. Am. Chem. Soc., 2014, 136,
16270–16276.
76 W. Zhang, L. Sun, J. M. V. Nsanzimana and X. Wang, Adv.
Mater.,2018, 30, 1705523.
77 H. Jin, C. Guo, X. Liu, J. Liu, A. Vasileff, Y. Jiao, Y. Zheng
and S.-Z. Qiao, Chem. Rev., 2018, 118, 6337–6408.
78 H. Jin, H. Yu, H. Li, K. Davey, T. Song, U. Paik and
S.-Z. Qiao, Angew. Chem., Int. Ed., 2022, 61, e202203850.
79 Y. Wu, D. Li, C.-L. Wu, H. Y. Hwang and Y. Cui, Nat. Rev.
Mater., 2023, 8, 41–53.
80 A. K. Geim, Science, 2009, 324, 1530–1534.
81 Z. Chen, C. Liu, L. Sun and T. Wang, ACS Catal., 2022, 12,
8936–8975.
82 C. N. R. Rao, A. K. Sood, K. S. Subrahmanyam and
A. Govindaraj, Angew. Chem., Int. Ed., 2009, 48, 7752–7777.
83 J. Abraham, K. S. Vasu, C. D. Williams, K. Gopinadhan,
Y. Su, C. T. Cherian, J. Dix, E. Prestat, S. J. Haigh,
I. V. Grigorieva, P. Carbone, A. K. Geim and R. R. Nair,
Nat. Nanotechnol., 2017, 12, 546–550.
84 A. W. Robertson, G.-D. Lee, K. He, C. Gong, Q. Chen,
E. Yoon, A. I. Kirkland and J. H. Warner, ACS Nano,
2015, 9, 11599–11607.
85 X.Chen,S.Zhang,D.Hou,H.Duan,B.Deng,Z.Zeng,B.Liu,
L.Sun,R.Song,J.Du,P.Gao,H.Peng,Z.LiuandL.Wang,
ACS Appl. Mater. Interfaces, 2021, 13, 29926–29935.
86 A. VahidMohammadi, J. Rosen and Y. Gogotsi, Science,
2021, 372, eabf1581.
87 M. Naguib, M. Kurtoglu, V. Presser, J. Lu, J. Niu, M. Heon,
L. Hultman, Y. Gogotsi and M. W. Barsoum, Adv. Mater.,
2011, 23, 4248–4253.
88 Y. Sun, S. Li, Y. Zhuang, G. Liu, W. Xing and W. Jing,
J. Membr. Sci., 2019, 591, 117350.
89 C. Tang, Y. Min, C. Chen, W. Xu and L. Xu, Nano Lett.,
2019, 19, 5577–5586.
90 J. Luo, J. Zheng, J. Nai, C. Jin, H. Yuan, O. Sheng, Y. Liu,
R. Fang, W. Zhang, H. Huang, Y. Gan, Y. Xia, C. Liang,
J. Zhang, W. Li and X. Tao, Adv. Funct. Mater., 2019,
29, 1808107.
91 Z. Wang, X. Zheng, A. Chen, Y. Han, L. Wei and J. Li, ACS
mater. lett., 2022, 4, 1436–1445.
92 W. Zhao, Y. Lei, Y. Zhu, Q. Wang, F. Zhang, X. Dong and
H. N. Alshareef, Nano Energy, 2021, 86, 106120.
93 Q. Yun, L. Li, Z. Hu, Q. Lu, B. Chen and H. Zhang, Adv.
Mater.,2020, 32, 1903826.
94 X. Chia and M. Pumera, Nat. Catal., 2018, 1, 909–921.
95 D. Monga, S. Sharma, N. P. Shetti, S. Basu, K. R. Reddy and
T. M. Aminabhavi, Mater. Today Chem., 2021, 19, 100399.
96 M. Chhowalla, H. S. Shin, G. Eda, L.-J. Li, K. P. Loh and
H. Zhang, Nat. Chem., 2013, 5, 263–275.
97 M. K. Aydinol, A. F. Kohan, G. Ceder, K. Cho and
J. Joannopoulos, Phys. Rev. B, 1997, 56, 1354–1365.
98 C. Zhu, D. Gao, J. Ding, D. Chao and J. Wang, Chem. Soc.
Rev., 2018, 47, 4332–4356.
99 Q. Yun, Q. Lu, X. Zhang, C. Tan and H. Zhang, Angew.
Chem., Int. Ed., 2018, 57, 626–646.
100 X. Yin, C. S. Tang, Y. Zheng, J. Gao, J. Wu, H. Zhang,
M. Chhowalla, W. Chen and A. T. S. Wee, Chem. Soc. Rev.,
2021, 50, 10087–10115.
101 M. Faizan, S. Hussain, D. Vikraman, B. Ali, H.-S. Kim,
J. Jung and K.-W. Nam, J. Mater. Res. Technol, 2021, 14,
2382–2393.
102 L. Shi and T. Zhao, J. Mater. Chem. A, 2017, 5, 3735–3758.
103 C. Vaalma, D. Buchholz, M. Weil and S. Passerini, Nat. Rev.
Mater., 2018, 3, 18013.
104 Y. Wang, X. Yu, S. Xu, J. Bai, R. Xiao, Y.-S. Hu, H. Li,
X.-Q. Yang, L. Chen and X. Huang, Nat. Commun., 2013,
4, 2365.
105 J.-Y. Li, H.-Y. Hu, L.-F. Zhou, H.-W. Li, Y.-J. Lei, W.-H. Lai,
Y.-M. Fan, Y.-F. Zhu, G. Peleckis, S.-Q. Chen, W.-K. Pang,
J. Peng, J.-Z. Wang, S.-X. Dou, S.-L. Chou and Y. Xiao, Adv.
Funct. Mater., 2023, 33, 2213215.
106 Y. Xiao, H.-R. Wang, H.-Y. Hu, Y.-F. Zhu, S. Li, J.-Y. Li,
X.-W. Wu and S.-L. Chou, Adv. Mater., 2022, 34, 2202695.
Chem Soc Rev Review Article
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2024
107 F. Wei, Q. Zhang, P. Zhang, W. Tian, K. Dai, L. Zhang,
J. Mao and G. Shao, J. Electrochem. Soc., 2021, 168, 050524.
108 J. Zhang, J.-B. Kim, J. Zhang, G.-H. Lee, M. Chen,
V. W.-H. Lau, K. Zhang, S. Lee, C.-L. Chen, T.-Y. Jeon,
Y.-W. Kwon and Y.-M. Kang, J. Am. Chem. Soc., 2022, 144,
7929–7938.
109 Z.-J. Zheng, H. Ye and Z.-P. Guo, Energy Environ. Sci., 2021,
14, 1835–1853.
110 J. Peng, W. Zhang, Q. Liu, J. Wang, S. Chou, H. Liu and
S. Dou, Adv. Mater., 2022, 34, 2108384.
111 H. Zhang, Y. Liu, J. Zhao, X. Peng, Y. Ren, X. Wei, Y. Song,
Z. Cao and Q. Wan, ChemistrySelect, 2022, 7, e202200785.
112 M. G. Moustafa and A. M. Aboraia, Ceram. Int., 2023, 49,
23068–23074.
113 S. Zheng, Z. Li, L. Chen, Y. Huang, J. Shi, S. Wang, Y. Liu,
Y. Liu, Y.-P. Cai and Q. Zheng, ACS mater. lett., 2023, 5,
1136–1144.
114 Y.-W. Song, P. Shi, B.-Q. Li, X. Chen, C.-X. Zhao,
W.-J. Chen, X.-Q. Zhang, X. Chen and Q. Zhang, Matter,
2021, 4, 253–264.
115 D. Zhu, G. Xu, M. Barnes, Y. Li, C.-P. Tseng, Z. Zhang,
J.-J. Zhang, Y. Zhu, S. Khalil, M. M. Rahman, R. Verduzco
and P. M. Ajayan, Adv. Funct. Mater., 2021, 31, 2100505.
116 B. Hu, J. Xu, Z. Fan, C. Xu, S. Han, J. Zhang, L. Ma, B. Ding,
Z. Zhuang, Q. Kang and X. Zhang, Adv. Energy Mater., 2023,
13, 2203540.
117 H. M. El-Kaderi, J. R. Hunt, J. L. Mendoza-Corte
´s,
A. P. Co
ˆte
´, R. E. Taylor, M. O’Keeffe and O. M. Yaghi,
Science, 2007, 316, 268–272.
118 M. E. Briggs and A. I. Cooper, Chem. Mater., 2017, 29,
149–157.
119 S. B. Kalidindi, S. Nayak, M. E. Briggs, S. Jansat,
A. P. Katsoulidis, G. J. Miller, J. E. Warren, D. Antypov,
F. Cora
`, B. Slater, M. R. Prestly, C. Martı
´-Gastaldo and
M. J. Rosseinsky, Angew. Chem., Int. Ed., 2015, 54, 221–226.
120 T. Hasell and A. I. Cooper, Nat. Rev. Mater., 2016, 1, 16053.
121 Y. Cheng, Y. Ying, S. Japip, S.-D. Jiang, T.-S. Chung,
S. Zhang and D. Zhao, Adv. Mater., 2018, 30,1802401.
122 X. Yang, Z. Ullah, J. F. Stoddart and C. T. Yavuz, Chem. Rev.,
2023, 123, 4602–4634.
123 K. Su, W. Wang, S. Du, C. Ji, M. Zhou and D. Yuan, J. Am.
Chem. Soc., 2020, 142, 18060–18072.
124 Y. Geng, T. Zhang, T. Xu, W. Mao, D. Li, K. Dai, J. Zhang
and G. Ai, Energy Storage Mater., 2022, 49, 67–76.
125 J. Ren, H. Zhu, Y. Fang, W. Li, S. Lan, S. Wei, Z. Yin,
Y. Tang, Y. Ren and Q. Liu, Carbon Neutraliz., 2023, 2,
339–377.
126 M. Chen, W. Hua, J. Xiao, J. Zhang, V. W.-H. Lau, M. Park,
G.-H. Lee, S. Lee, W. Wang, J. Peng, L. Fang, L. Zhou,
C.-K. Chang, Y. Yamauchi, S. Chou and Y.-M. Kang, J. Am.
Chem. Soc., 2021, 143, 18091–18102.
127 M. C. Orilall and U. Wiesner, Chem. Soc. Rev., 2011, 40,
520–535.
128 D. B. Mitzi, Chem. Mater., 2001, 13, 3283–3298.
129 D. P. Dubal, O. Ayyad, V. Ruiz and P. Go
´mez-Romero,
Chem. Soc. Rev., 2015, 44, 1777–1790.
130 E. Pomerantseva and Y. Gogotsi, Nat. Energy., 2017,
2, 17089.
131 P. Xiong, F. Zhang, X. Zhang, S. Wang, H. Liu, B. Sun,
J. Zhang, Y. Sun, R. Ma, Y. Bando, C. Zhou, Z. Liu, T. Sasaki
and G. Wang, Nat. Commun., 2020, 11, 3297.
132 P. Du, L. Cao, B. Zhang, C. Wang, Z. Xiao, J. Zhang,
D. Wang and X. Ou, Renew. Sustain. Energy Rev., 2021,
151, 111640.
133 T. Liu, Y. Zhao, L. Gao and J. Ni, Sci. Rep., 2015, 5, 9307.
134 F. Peng, L. Yu, S. Yuan, X.-Z. Liao, J. Wen, G. Tan, F. Feng
and Z.-F. Ma, ACS Appl. Mater. Interfaces, 2019, 11,
37685–37692.
135 B. Ni and X. Wang, Chem. Sci., 2016, 7, 3978–3991.
136 K. Mu
¨llen, Angew. Chem., Int. Ed., 2015, 54, 10040–10042.
137 I. Borukhov, D. Andelman and H. Orland, Phys. Rev. Lett.,
1997, 79, 435–438.
138 S. Kondrat and A. Kornyshev, J. Phys. Chem. C,2013, 117,
12399–12406.
139 K. Zhang, W. Liu, Y. Gao, X. Wang, Z. Chen, R. Ning,
W. Yu, R. Li, L. Li, X. Li, K. Yuan, L. Ma, N. Li, C. Shen,
W. Huang, K. Xie and K. P. Loh, Adv. Mater., 2021,
33, 2006323.
140 J. Yang, B. Yin, Y. Sun, H. Pan, W. Sun, B. Jia, S. Zhang and
T. Ma, Nano-Micro Lett., 2022, 14, 42.
141 Z. Hong, Y. Zhen, Y. Ruan, M. Kang, K. Zhou, J. M. Zhang,
Z. Huang and M. Wei, Adv. Mater., 2018, 30, 1802035.
142 W. Ye, X. Li, B. Zhang, W. Liu, Y. Cheng, X. Fan, H. Zhang,
Y. Liu, Q. Dong and M.-S. Wang, Adv. Mater., 2023,
35, 2210447.
143 H. Zhang, X. Li, J. Hou, L. Jiang and H. Wang, Chem. Soc.
Rev., 2022, 51, 2224–2254.
144 L. Wang, M. S. H. Boutilier, P. R. Kidambi, D. Jang,
N. G. Hadjiconstantinou and R. Karnik, Nat. Nanotechnol.,
2017, 12, 509–522.
145 A. R. Koltonow and J. Huang, Science, 2016, 351,
1395–1396.
146 C. L. Ritt, M. Liu, T. A. Pham, R. Epsztein, H. J. Kulik and
M. Elimelech, Sci. Adv., 2022, 8, eabl5771.
147 G. Cai, Y. Yin, D. Xia, A. A. Chen, J. Holoubek, J. Scharf,
Y. Yang, K. H. Koh, M. Li, D. M. Davies, M. Mayer,
T. H. Han, Y. S. Meng, T. A. Pascal and Z. Chen, Nat.
Commun., 2021, 12, 3395.
148 C. Cheng, G. Jiang, C. J. Garvey, Y. Wang, G. P. Simon,
J. Z. Liu and D. Li, Sci. Adv., 2016, 2, e1501272.
149 Y. Xu, L. Gao, L. Shen, Q. Liu, Y. Zhu, Q. Liu, L. Li, X. Kong,
Y. Lu and H. B. Wu, Matter, 2020, 3, 1685–1700.
150 M. Zhang, K. Guan, Y. Ji, G. Liu, W. Jin and N. Xu, Nat.
Commun., 2019, 10, 1253.
151 J. Jas
ˇı
´k, A. Fortunelli and S
ˇ. Vajda, Phys. Chem. Chem. Phys.,
2022, 24, 12083–12115.
152 Q. Zeng, L. Xu, G. Li, Q. Zhang, S. Guo, H. Lu, L. Xie,
J. Yang, J. Weng, C. Zheng and S. Huang, Adv. Funct.
Mater., 2023, 2304619.
153 S. Xin, L. Gu, N.-H. Zhao, Y.-X. Yin, L.-J. Zhou, Y.-G.
Guo and L.-J. Wan, J. Am. Chem. Soc., 2012, 134,
18510–18513.
Review Article Chem Soc Rev
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
This journal is © The Royal Society of Chemistry 2024 Chem. Soc. Rev.
154 C. Wang, Y. Yang, X. Liu, H. Zhong, H. Xu, Z. Xu, H.
Shao and F. Ding, ACS Appl. Mater. Interfaces, 2017, 9,
13694–13702.
155 Q. Zhang, B. Yue, C. Shao, H. Shao, L. Li, X. Dong, J. Wang
and W. Yu, J. Chem. Eng, 2022, 443, 136479.
156 A. Taloni, M. Vodret, G. Costantini and S. Zapperi, Nat.
Rev. Mater, 2018, 3, 211–224.
157 J. F. Go
´mez-Corte
´s, M. L. No
´,I.Lo
´pez-Ferren
˜o, J.
Herna
´ndez-Saz, S. I. Molina, A. Chuvilin and J. M. San
Juan, Nat. Nanotechnol., 2017, 12, 790–796.
158 G. L. Gregory, H. Gao, B. Liu, X. Gao, G. J. Rees, M. Pasta,
P. G. Bruce and C. K. Williams, J. Am. Chem. Soc., 2022,
144, 17477–17486.
159 S. Rajendran, Z. Tang, A. George, A. Cannon, C. Neumann,
A. Sawas, E. Ryan, A. Turchanin and L. M. R. Arava, Adv.
Energy Mater., 2021, 11, 2100666.
160 M. Majumder, N. Chopra, R. Andrews and B. J. Hinds,
Nature, 2005, 438, 44.
161 J. Shen, G. Liu, Y. Han and W. Jin, Nat. Rev. Mater, 2021, 6,
294–312.
162 R. Epsztein, R. M. DuChanois, C. L. Ritt, A. Noy and
M. Elimelech, Nat. Nanotechnol., 2020, 15, 426–436.
163 Z. Zhu, D. Wang, Y. Tian and L. Jiang, J. Am. Chem. Soc.,
2019, 141, 8658–8669.
164 D. Pan, J. Zhang, Z. Li and M. Wu, Adv. Mater., 2010, 22,
734–738.
165 S. Levasseur, M. Me
´ne
´trier, Y. Shao-Horn, L. Gautier,
A. Audemer, G. Demazeau, A. Largeteau and C. Delmas,
Chem. Mater., 2003, 15, 348–354.
166 J.-M. Lim, D. Kim, Y.-G. Lim, M.-S. Park, Y.-J. Kim, M. Cho
and K. Cho, ChemElectroChem, 2016, 3, 943–949.
167 Z. Wang, L. Liu, L. Chen and X. Huang, Solid State Ion.,
2002, 148, 335–342.
168 F. Ning, S. Li, B. Xu and C. Ouyang, Solid State Ion., 2014,
263, 46–48.
169 G. Qi, J. Hu, M. Balogh, L. Wang, D. Darbar and W. Li,
Electrochem, 2023, 4, 21–30.
170 H. Xie, G. Hu, K. Du, Z. Peng and Y. Cao, J. Alloys Compd.,
2016, 666, 84–87.
171 H. S. Ko, J. H. Kim, J. Wang and J. D. Lee, J. Power Sources,
2017, 372, 107–115.
172 S. Lee, Y. Cho, H.-K. Song, K. T. Lee and J. Cho, Angew.
Chem., Int. Ed., 2012, 51, 8748–8752.
173 P. J. Phillips, J. Baren
˜o, Y. Li, D. P. Abraham and R. F. Klie,
Adv. Energy Mater., 2015, 5, 1501252.
174 K. Du, H. Xie, G. Hu, Z. Peng, Y. Cao and F. Yu, ACS Appl.
Mater. Interfaces, 2016, 8, 17713–17720.
175 G. Dai, H. Du, S. Wang, J. Cao, M. Yu, Y. Chen,
Y. Tang, A. Li and Y. Chen, RSC Adv., 2016, 6,
100841–100848.
176 Y. Liu, Z. Yang, J. Zhong, J. Li, R. Li, Y. Yu and F. Kang, ACS
Nano, 2019, 13, 11891–11900.
177 C. Ma, X.-L. Li, X.-Y. Yue, J. Bao, R.-J. Luo and Y.-N. Zhou,
J. Chem. Eng., 2022, 432, 134305.
178 M. Si, D. Wang, R. Zhao, D. Pan, C. Zhang, C. Yu, X. Lu,
H. Zhao and Y. Bai, Adv. Sci., 2020, 7, 1902538.
179 B. J. Hwang, R. Santhanam and C. H. Chen, J. Power
Sources, 2003, 114, 244–252.
180 W. Li and J. C. Currie, J. Electrochem. Soc., 1997, 144, 2773.
181 R. V. Chebiam, F. Prado and A. Manthiram, J. Solid State
Chem., 2002, 163, 5–9.
182 F. Wu, M. Zhang, Y. Bai, X. Wang, R. Dong and C. Wu, ACS
Appl. Mater. Interfaces, 2019, 11, 12554–12561.
183 V.-C. Ho, H. An, M. Hong, S. Lee, J. Kim, M. B. Park and
J. Mun, Energy Technol., 2021, 9, 2000800.
184 F. Schipper, M. Dixit, D. Kovacheva, M. Talianker, O. Haik,
J. Grinblat, E. M. Erickson, C. Ghanty, D. T. Major,
B. Markovsky and D. Aurbach, J. Mater. Chem. A, 2016, 4,
16073–16084.
185 N. Sun, Z. Guan, Y. Liu, Y. Cao, Q. Zhu, H. Liu, Z. Wang,
P. Zhang and B. Xu, Adv. Energy Mater., 2019, 9, 1901351.
186 Y. Zhao, J. Liu, S. Wang, R. Ji, Q. Xia, Z. Ding, W. Wei,
Y. Liu, P. Wang and D. G. Ivey, Adv. Funct. Mater., 2016, 26,
4760–4767.
187 L. Ma, L. Mao, X. Zhao, J. Lu, F. Zhang, P. Ding, L. Chen
and F. Lian, ChemElectroChem, 2017, 4, 3068–3074.
188 Q. Hu, Y. He, D. Ren, Y. Song, Y. Wu, H. Liang, J. Gao,
G. Xu, J. Cai, T. Li, H. Xu, L. Wang, Z. Chen and X. He,
Nano Energy, 2022, 96, 107123.
189 W. Li, X. Liu, H. Celio, P. Smith, A. Dolocan, M. Chi and
A. Manthiram, Adv. Energy Mater., 2018, 8, 1703154.
190 R. Va
¨li, J. Aruva
¨li, M. Ha
¨rmas, A. Ja
¨nes and E. Lust,
Batteries, 2019, 5, 56.
191 F. Xin, H. Zhou, Y. Zong, M. Zuba, Y. Chen, N. A. Chernova,
J. Bai, B. Pei, A. Goel, J. Rana, F. Wang, K. An, L. F. J. Piper,
G. Zhou and M. S. Whittingham, ACS Energy Lett., 2021, 6,
1377–1382.
192 S. Liu, X. Chen, J. Zhao, J. Su, C. Zhang, T. Huang, J. Wu
and A. Yu, J. Power Sources, 2018, 374, 149–157.
193 J. Li, M. Zhang, D. Zhang, Y. Yan and Z. Li, J. Chem. Eng.,
2020, 402, 126195.
194 C. Matei Ghimbeu, B. Zhang, A. Martinez de Yuso, B. Re
´ty
and J.-M. Tarascon, Carbon, 2019, 153, 634–647.
195 A. Gomez-Martin, F. Reissig, L. Frankenstein,
M. Heidbu
¨chel, M. Winter, T. Placke and R. Schmuch,
Adv. Energy Mater., 2022, 12, 2103045.
196 A. Windmu
¨ller, C.-L. Tsai, S. Mo
¨ller, M. Balski, Y. J. Sohn,
S. Uhlenbruck and O. Guillon, J. Power Sources, 2017, 341,
122–129.
197 A. Eftekhari, J. Power Sources, 2003, 124, 182–190.
198 S. Mukerjee, X. Q. Yang, X. Sun, S. J. Lee, J. McBreen and
Y. Ein-Eli, Electrochim. Acta, 2004, 49, 3373–3382.
199 J. Xu, Z. Han, K. Jiang, P. Bai, Y. Liang, X. Zhang, P. Wang,
S. Guo and H. Zhou, Small, 2020, 16, 1904388.
200 P.-F. Wang, Y. You, Y.-X. Yin and Y.-G. Guo, J. Mater. Chem.
A, 2016, 4, 17660–17664.
201 P.-F. Wang, H.-R. Yao, X.-Y. Liu, J.-N. Zhang, L. Gu,
X.-Q. Yu, Y.-X. Yin and Y.-G. Guo, Adv. Mater., 2017,
29, 1700210.
202 S. Zheng, G. Zhong, M. J. McDonald, Z. Gong, R. Liu,
W. Wen, C. Yang and Y. Yang, J. Mater. Chem. A, 2016, 4,
9054–9062.
Chem Soc Rev Review Article
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2024
203 Y. Zhang, N. Zhang, W. Chen, Z. Rao, J. Wu, L. Xue and
W. Zhang, Energy Technol., 2019, 7, 1900779.
204 L. Zhang, T. Yuan, L. Soule, H. Sun, Y. Pang, J. Yang and
S. Zheng, ACS Appl. Energy Mater., 2020, 3, 3770–3778.
205 Y. Shen, S. Sun, M. Yang and X. Zhao, J. Alloys Compd.,
2019, 784, 1290–1296.
206 L. Sun, Y. Xie, X.-Z. Liao, H. Wang, G. Tan, Z. Chen, Y. Ren,
J. Gim, W. Tang, Y.-S. He, K. Amine and Z.-F. Ma, Small,
2018, 14, 1704523.
207 S.-M. Oh, S.-T. Myung, J.-Y. Hwang, B. Scrosati, K. Amine
and Y.-K. Sun, Chem. Mater., 2014, 26, 6165–6171.
208 Y. You, S. Xin, H. Y. Asl, W. Li, P.-F. Wang, Y.-G. Guo and
A. Manthiram, Chem, 2018, 4, 2124–2139.
209 J.-L. Yue, Y.-N. Zhou, X. Yu, S.-M. Bak, X.-Q. Yang and
Z.-W. Fu, J. Mater. Chem. A, 2015, 3, 23261–23267.
210 C. Zhang, R. Gao, L. Zheng, Y. Hao and X. Liu, ACS Appl.
Mater. Interfaces, 2018, 10, 10819–10827.
211 M.-H. Cao, Y. Wang, Z. Shadike, J.-L. Yue, E. Hu, S.-M. Bak,
Y.-N. Zhou, X.-Q. Yang and Z.-W. Fu, J. Mater. Chem. A,
2017, 5, 5442–5448.
212 L. Mu, S. Xu, Y. Li, Y.-S. Hu, H. Li, L. Chen and X. Huang,
Adv. Mater., 2015, 27, 6928–6933.
213 C. G. Hawkins and L. Whittaker-Brooks, ACS Appl. Nano
Mater., 2018, 1, 851–859.
214 C. Vaalma, G. A. Giffin, D. Buchholz and S. Passerini,
J. Electrochem. Soc., 2016, 163, A1295.
215 H. Kim, J. C. Kim, S.-H. Bo, T. Shi, D.-H. Kwon and
G. Ceder, Adv. Energy Mater., 2017, 7, 1700098.
216 X. Wang, X. Xu, C. Niu, J. Meng, M. Huang, X. Liu, Z. Liu
and L. Mai, Nano Lett., 2017, 17, 544–550.
217 J. J. Ding, Y. N. Zhou, Q. Sun, X. Q. Yu, X. Q. Yang and
Z. W. Fu, Electrochim. Acta, 2013, 87, 388–393.
218 Y. Fang, X.-Y. Yu and X. W. Lou, Angew. Chem., Int. Ed.,
2017, 56, 5801–5805.
219 P. Vassilaras, X. Ma, X. Li and G. Ceder, J. Electrochem. Soc.,
2013, 160, A207.
220 N. Yabuuchi, H. Yoshida and S. Komaba, Electrochem.,
2012, 80, 716–719.
221 C.-Y. Yu, J.-S. Park, H.-G. Jung, K.-Y. Chung, D. Aurbach,
Y.-K. Sun and S.-T. Myung, Energy Environ. Sci., 2015, 8,
2019–2026.
222 D. Hamani, M. Ati, J.-M. Tarascon and P. Rozier, Electro-
chem. Commun., 2011, 13, 938–941.
223 X. Wang, P. Hu, C. Niu, J. Meng, X. Xu, X. Wei, C. Tang,
W. Luo, L. Zhou, Q. An and L. Mai, Nano Energy, 2017, 35,
71–78.
224 H. Zhang, K. Xi, K. Jiang, X. Zhang, Z. Liu, S. Guo and
H. Zhou, Chem. comm., 2019, 55, 7910–7913.
225 S. Guo, H. Yu, P. Liu, Y. Ren, T. Zhang, M. Chen,
M. Ishida and H. Zhou, Energy Environ. Sci., 2015, 8,
1237–1244.
226 Y. Wang, R. Xiao, Y.-S. Hu, M. Avdeev and L. Chen, Nat.
Commun., 2015, 6, 6954.
227 J. Xu, D. H. Lee, R. J. Cle
´ment, X. Yu, M. Leskes, A. J. Pell,
G. Pintacuda, X.-Q. Yang, C. P. Grey and Y. S. Meng, Chem.
Mater., 2014, 26, 1260–1269.
228 J. Yoshida, E. Guerin, M. Arnault,C.Constantin,B.Mortemard
de Boisse, D. Carlier, M. Guignard and C. Delmas,
J. Electrochem. Soc., 2014, 161, A1987.
229 Y. Hironaka, K. Kubota and S. Komaba, Chem. comm.,
2017, 53, 3693–3696.
230 C.-L. Liu, S.-H. Luo, H.-B. Huang, Y.-C. Zhai and Z.-W.
Wang, J. Chem. Eng., 2019, 356, 53–59.
231 H. Kim, D.-H. Seo, A. Urban, J. Lee, D.-H. Kwon, S.-H. Bo,
T. Shi, J. K. Papp, B. D. McCloskey and G. Ceder, Chem.
Mater., 2018, 30, 6532–6539.
232 L. Deng, X. Niu, G. Ma, Z. Yang, L. Zeng, Y. Zhu and L. Guo,
Adv. Funct. Mater., 2018, 28, 1800670.
233 J. U. Choi, J. Kim, J.-Y. Hwang, J. H. Jo, Y.-K. Sun and
S.-T. Myung, Nano Energy, 2019, 61, 284–294.
234 T. Deng, X. Fan, J. Chen, L. Chen, C. Luo, X. Zhou,
J. Yang, S. Zheng and C. Wang, Adv. Funct. Mater., 2018,
28, 1800219.
235 X. Zhang, Y. Yang, X. Qu, Z. Wei, G. Sun, K. Zheng, H. Yu
and F. Du, Adv. Funct. Mater., 2019, 29,1905679.
236 K. Lei, Z. Zhu, Z. Yin, P. Yan, F. Li and J. Chen, Chem, 2019,
5, 3220–3231.
237 S. Zhao, K. Yan, P. Munroe, B. Sun and G. Wang, Adv.
Energy Mater., 2019, 9, 1803757.
238 B. Peng, Y. Li, J. Gao, F. Zhang, J. Li and G. Zhang, J. Power
Sources, 2019, 437, 226913.
239 N. Naveen, W. B. Park, S. P. Singh, S. C. Han, D. Ahn,
K.-S. Sohn and M. Pyo, Small, 2018, 14, 1803495.
240 J.-Y. Hwang, J. Kim, T.-Y. Yu, S.-T. Myung and Y.-K. Sun,
Energy Environ. Sci., 2018, 11, 2821–2827.
241 N. Naveen, S. C. Han, S. P. Singh, D. Ahn, K.-S. Sohn and
M. Pyo, J. Power Sources, 2019, 430, 137–144.
242 Y. Yang, Z. Liu, L. Deng, L. Tan, X. Niu, M. M. S.
Sanad, L. Zeng, Z. Zhu and Y. Zhu, Chem. comm., 2019,
55, 14988–14991.
243 X. Niu, J. Qu, Y. Hong, L. Deng, R. Wang, M. Feng, J. Wang,
L. Zeng, Q. Zhang, L. Guo and Y. Zhu, J. Mater. Chem. A,
2021, 9, 13125–13134.
244 J. H. Jo, J.-Y. Hwang, J. U. Choi, H. J. Kim, Y.-K. Sun and
S.-T. Myung, J. Power Sources, 2019, 432, 24–29.
245 Y. Zhang, X. Niu, L. Tan, L. Deng, S. Jin, L. Zeng, H.
Xu and Y. Zhu, ACS Appl. Mater. Interfaces, 2020, 12,
9332–9340.
246 W. Li, Z. Bi, W. Zhang, J. Wang, R. Rajagopalan, Q. Wang,
D. Zhang, Z. Li, H. Wang and B. Wang, J. Mater. Chem. A,
2021, 9, 8221–8247.
247 C.-L. Liu, S.-H. Luo, H.-B. Huang, Y.-C. Zhai and Z.-W.
Wang, ChemElectroChem, 2019, 6, 2308–2315.
248 J. Weng, J. Duan, C. Sun, P. Liu, A. Li, P. Zhou and J. Zhou,
J. Chem. Eng., 2020, 392, 123649.
249 M. G. T. Nathan, N. Naveen, W. B. Park, K.-S. Sohn and
M. Pyo, J. Power Sources, 2019, 438, 226992.
250 J. U. Choi, Y. Ji Park, J. H. Jo, Y. H. Jung, D.-C.
Ahn, T.-Y. Jeon, K.-S. Lee, H. Kim, S. Lee, J. Kim and
S.-T. Myung, Energy Storage Mater, 2020, 27, 342–351.
251 P. Bai, K. Jiang, X. Zhang, J. Xu, S. Guo and H. Zhou, ACS
Appl. Mater. Interfaces, 2020, 12, 10490–10495.
Review Article Chem Soc Rev
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
This journal is © The Royal Society of Chemistry 2024 Chem. Soc. Rev.
252 H. V. Ramasamy, B. Senthilkumar, P. Barpanda and
Y.-S. Lee, J. Chem. Eng., 2019, 368, 235–243.
253 C. Liu, S. Luo, H. Huang, Z. Wang, A. Hao, Y. Zhai and
Z. Wang, Electrochem. Commun., 2017, 82, 150–154.
254 Z. A. Ghazi, X. He, A. M. Khattak, N. A. Khan, B. Liang,
A. Iqbal, J. Wang, H. Sin, L. Li and Z. Tang, Adv. Mater.,
2017, 29, 1606817.
255 Y. Ge, X. Yan, J. Liu, X. Zhang, J. Wang, X. He, R. Wang and
H. Xie, Electrochim. Acta, 2010, 55, 5886–5890.
256 H. Liu, S.-H. Luo, S.-X. Yan, Y.-F. Wang, Q. Wang, M.-Q. Li
and Y.-H. Zhang, J. Electroanal. Chem., 2019, 850, 113434.
257 X. Li, Z. Shao, K. Liu, Q. Zhao, G. Liu and B. Xu, J. Solid
State Electrochem., 2019, 23, 465–473.
258 T. A. Wani and G. Suresh, J. Energy Storage, 2021,
44, 103307.
259 I. Seo, B. Senthilkumar, K.-H. Kim, J.-K. Kim, Y. Kim and
J.-H. Ahn, J. Power Sources, 2016, 320, 59–67.
260 M. Zhang, N. Garcia-Araez and A. L. Hector, J. Mater. Chem.
A, 2018, 6, 14483–14517.
261 Y. Xiao, P.-F. Wang, Y.-X. Yin, Y.-F. Zhu, X. Yang,
X.-D. Zhang, Y. Wang, X.-D. Guo, B.-H. Zhong and
Y.-G. Guo, Adv. Energy Mater., 2018, 8, 1800492.
262 K. Vijaya Babu, L. Seeta Devi, V. Veeraiah and K. Anand,
J. Asian Ceram. Soc., 2016, 4, 269–276.
263 Q. Ni, L. Zheng, Y. Bai, T. Liu, H. Ren, H. Xu, C. Wu and
J. Lu, ACS Energy Lett., 2020, 5, 1763–1770.
264 X. Hou, J. Liang, T. Zhang, Y. Li, S. Tang, H. Sun, J. Zhang
and H. Xie, J. Phys. Chem. C, 2017, 121, 22656–22664.
265 T. Yamada, N. Zettsu, H.-M. Kim, Y. Hagano, N. Handa,
K. Yubuta and K. Teshima, Cryst. Growth Des., 2018, 18,
6777–6785.
266 S. S. Fedotov, A. A. Kabanov, N. A. Kabanova, V. A. Blatov,
A. Zhugayevych, A. M. Abakumov, N. R. Khasanova and
E. V. Antipov, J. Phys. Chem. C, 2017, 121, 3194–3202.
267 M. Sun, G. Rousse, D. D. Corte, M. Saubane
`re, M.-L.
Doublet and J.-M. Tarascon, Chem. Mater., 2016, 28,
1607–1610.
268 C. Frayret, A. Villesuzanne, N. Spaldin, E. Bousquet,
J.-N. Chotard, N. Recham and J.-M. Tarascon, Phys. Chem.
Chem. Phys., 2010, 12, 15512–15522.
269 T. Mueller, G. Hautier, A. Jain and G. Ceder, Chem. Mater.,
2011, 23, 3854–3862.
270 W. Wen, X. Wang, S. Chen, H. Shu and X. Yang, J. Power
Sources, 2015, 281, 85–93.
271 M. H. Nasir, N. K. Janjua and J. Santoki, J. Electrochem.
Soc., 2020, 167, 130526.
272 X. Wu, Z. Gong, S. Tan and Y. Yang, J. Power Sources, 2012,
220, 122–129.
273 N. Membren
˜o, K. Park, J. B. Goodenough and
K. J. Stevenson, Chem. Mater., 2015, 27, 3332–3340.
274 Y. Liu, Z. Wu, S. Indris, W. Hua, N. P. M. Casati, A. Tayal,
M. S. D. Darma, G. Wang, Y. Liu, C. Wu, Y. Xiao, B. Zhong
and X. Guo, Nano Energy, 2021, 79, 105417.
275 M. Hadouchi, N. Yaqoob, P. Kaghazchi, M. Tang, J. Liu,
P. Sang, Y. Fu, Y. Huang and J. Ma, Energy Storage Mater.,
2021, 35, 192–202.
276 S.-M. Oh, S.-T. Myung, J. Hassoun, B. Scrosati and
Y.-K. Sun, Electrochemistry Communications, 2012, 22,
149–152.
277 Y. Liu, N. Zhang, F. Wang, X. Liu, L. Jiao and L.-Z. Fan, Adv.
Funct. Mater., 2018, 28, 1801917.
278 Y. Liu, Y. Xu, X. Han, C. Pellegrinelli, Y. Zhu, H. Zhu,
J. Wan, A. C. Chung, O. Vaaland, C. Wang and L. Hu, Nano
Lett., 2012, 12, 5664–5668.
279 G. He, W. H. Kan and A. Manthiram, Chem. Mater., 2016,
28, 682–688.
280 Y. Fang, Q. Liu, L. Xiao, Y. Rong, Y. Liu, Z. Chen, X. Ai,
Y. Cao, H. Yang, J. Xie, C. Sun, X. Zhang, B. Aoun, X. Xing,
X. Xiao and Y. Ren, Chem, 2018, 4, 1167–1180.
281 J. Kim, G. Yoon, H. Kim, Y.-U. Park and K. Kang, Chem.
Mater., 2018, 30, 3683–3689.
282 K. Saravanan, C. W. Mason, A. Rudola, K. H. Wong and
P. Balaya, Adv. Energy Mater., 2013, 3, 444–450.
283 R. Rajagopalan, B. Chen, Z. Zhang, X.-L. Wu, Y. Du,
Y.Huang,B.Li,Y.Zong,J.Wang,G.-H.Nam,M.Sindoro,
S. X. Dou, H. K. Liu and H. Zhang, Adv. Mater., 2017,
29, 1605694.
284 K. Kawai, W. Zhao, S.-I. Nishimura and A. Yamada, ACS
Appl. Energy Mater, 2018, 1, 928–931.
285 D. Wang, X. Bie, Q. Fu, D. Dixon, N. Bramnik, Y.-S. Hu,
F. Fauth, Y. Wei, H. Ehrenberg, G. Chen and F. Du, Nat.
Commun., 2017, 8, 15888.
286 W. Zhou, L. Xue, X. Lu
¨, H. Gao, Y. Li, S. Xin, G. Fu, Z. Cui,
Y. Zhu and J. B. Goodenough, Nano Lett., 2016, 16,
7836–7841.
287 H. Gao, I. D. Seymour, S. Xin, L. Xue, G. Henkelman and
J. B. Goodenough, J. Am. Chem. Soc., 2018, 140, 18192–18199.
288 H. Li, M. Xu, C. Gao, W. Zhang, Z. Zhang, Y. Lai and L. Jiao,
Energy Storage Mater., 2020, 26, 325–333.
289 C. S. Park, H. Kim, R. A. Shakoor, E. Yang, S. Y. Lim,
R. Kahraman, Y. Jung and J. W. Choi, J. Am. Chem. Soc.,
2013, 135, 2787–2792.
290 P. Barpanda, J. Lu, T. Ye, M. Kajiyama, S.-C. Chung,
N. Yabuuchi, S. Komaba and A. Yamada, RSC Adv., 2013,
3, 3857–3860.
291 B. Lin, S. Zhang and C. Deng, J. Mater. Chem. A, 2016, 4,
2550–2559.
292 J. Kim, I. Park, H. Kim, K.-Y. Park, Y.-U. Park and K. Kang,
Adv. Energy Mater., 2016, 6, 1502147.
293 P. Barpanda, G. Liu, M. Avdeev and A. Yamada, ChemElec-
troChem, 2014, 1, 1488–1491.
294 Y. Kee, N. Dimov, A. Staikov, P. Barpanda, Y.-C. Lu,
K. Minami and S. Okada, RSC Adv., 2015, 5, 64991–64996.
295 M. Zarrabeitia, M. Ja
´uregui, N. Sharma, J. C. Pramudita
and M. Casas-Cabanas, Chem. Mater., 2019, 31, 5152–5159.
296 H. Zhang, I. Hasa, D. Buchholz, B. Qin, D. Geiger, S. Jeong,
U. Kaiser and S. Passerini, NPG Asia Mater., 2017, 9,
e370–e370.
297 W. Fang, Z. An, J. Xu, H. Zhao and J. Zhang, RSC Adv., 2018,
8, 21224–21228.
298 T. Jin, Y. Liu, Y. Li, K. Cao, X. Wang and L. Jiao, Adv. Energy
Mater., 2017, 7, 1700087.
Chem Soc Rev Review Article
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2024
299 J.-Z. Guo, P.-F. Wang, X.-L. Wu, X.-H. Zhang, Q. Yan,
H. Chen, J.-P. Zhang and Y.-G. Guo, Adv. Mater., 2017,
29, 1701968.
300 R. Tripathi, S. M. Wood, M. S. Islam and L. F. Nazar, Energy
Environ. Sci., 2013, 6, 2257–2264.
301 H. Zou, S. Li, X. Wu, M. J. McDonald and Y. Yang, ECS
Electrochem. Lett., 2015, 4, A53.
302 W. Pan, W. Guan, S. Liu, B. B. Xu, C. Liang, H. Pan,
M. Yan and Y. Jiang, J. Mater. Chem. A, 2019, 7,
13197–13204.
303 P. Barpanda, G. Oyama, C. D. Ling and A. Yamada, Chem.
Mater., 2014, 26, 1297–1299.
304 M. Chen, D. Cortie, Z. Hu, H. Jin, S. Wang, Q. Gu, W. Hua,
E. Wang, W. Lai, L. Chen, S.-L. Chou, X.-L. Wang and
S.-X. Dou, Adv. Energy Mater., 2018, 8, 1800944.
305 A. Gon
˜i, A. Iturrondobeitia, I. Gil de Muro, L. Lezama and
T. Rojo, J. Power Sources, 2017, 369, 95–102.
306 J. Lin, Y. Xu, J. Wang, B. Zhang, C. Wang, S. He and J. Wu,
Chem. Eng. J., 2019, 373, 78–85.
307 P. Singh, K. Shiva, H. Celio and J. B. Goodenough, Energy
Environ. Sci., 2015, 8, 3000–3005.
308 M. Law, V. Ramar and P. Balaya, J. Power Sources, 2017,
359, 277–284.
309 K. Kaliyappan and Z. Chen, Electrochim. Acta, 2018, 283,
1384–1389.
310 J. C. Treacher, S. M. Wood, M. S. Islam and E. Kendrick,
Phys. Chem. Chem. Phys., 2016, 18, 32744–32752.
311 M. Chen, W. Hua, J. Xiao, D. Cortie, W. Chen, E. Wang,
Z. Hu, Q. Gu, X. Wang, S. Indris, S.-L. Chou and S.-X. Dou,
Nat. Commun., 2019, 10, 1480.
312 J. Han, Y. Niu, S.-J. Bao, Y.-N. Yu, S.-Y. Lu and M. Xu, Chem.
Commun., 2016, 52, 11661–11664.
313 N. Recham, G. Rousse, M. T. Sougrati, J.-N. Chotard,
C. Frayret, S. Mariyappan, B. C. Melot, J.-C. Jumas and
J.-M. Tarascon, Chem. Mater., 2012, 24, 4363–4370.
314 J. Han, G.-N. Li, F. Liu, M. Wang, Y. Zhang, L. Hu, C. Dai
and M. Xu, Chem. comm., 2017, 53, 1805–1808.
315 X. Lin, J. Huang, H. Tan, J. Huang and B. Zhang, Energy
Storage Mater., 2019, 16, 97–101.
316 J. Liao, Q. Hu, B. Che, X. Ding, F. Chen and C. Chen,
J. Mater. Chem. A, 2019, 7, 15244–15251.
317 N. Imanishi, T. Morikawa, J. Kondo, Y. Takeda, O.
Yamamoto, N. Kinugasa and T. Yamagishi, J. Power
Sources, 1999, 79, 215–219.
318 L. Shen, Z. Wang and L. Chen, Chem. Eur. J., 2014, 20,
12559–12562.
319 X. Wu, M. Shao, C. Wu, J. Qian, Y. Cao, X. Ai and H. Yang,
ACS Appl. Mater. Interfaces, 2016, 8, 23706–23712.
320 Z. Zhang, M. Avdeev, H. Chen, W. Yin, W. H. Kan and
G. He, Nat. Commun., 2022, 13, 7790.
321 Y. You, X.-L. Wu, Y.-X. Yin and Y.-G. Guo, J. Mater. Chem. A,
2013, 1, 14061–14065.
322 H. Lee, Y.-I. Kim, J.-K. Park and J. W. Choi, Chem. comm.,
2012, 48, 8416–8418.
323 J. Peng, M. Ou, H. Yi, X. Sun, Y. Zhang, B. Zhang, Y. Ding,
F. Wang, S. Gu, C. A. Lo
´pez, W. Zhang, Y. Liu, J. Fang, P.
Wei,Y.Li,L.Miao,J.Jiang,C.Fang,Q.Li,M.T.Ferna
´ndez-
´
az,J.A.Alonso,S.ChouandJ.Han,Energy Environ. Sci.,
2021, 14, 3130–3140.
324 W. Wang, Y. Gang, Z. Hu, Z. Yan, W. Li, Y. Li, Q.-F. Gu,
Z. Wang, S.-L. Chou, H.-K. Liu and S.-X. Dou, Nat. Com-
mun., 2020, 11, 980.
325 J. Peng, W. Zhang, J. Wang, L. Li, W. Lai, Q. Yang,
B. Zhang, X. Li, Y. Du, H. Liu, J. Wang, Z. Cheng,
L. Wang, S. Wang, J. Wang, S. Chou, H. Liu and S. Dou,
Adv. Energy Mater., 2021, 11, 2102356.
326 Y. Shang, X. Li, J. Song, S. Huang, Z. Yang, Z. J. Xu and
H. Y. Yang, Chem, 2020, 6, 1804–1818.
327 Y. Liu, Y. Qiao, W. Zhang, Z. Li, X. Ji, L. Miao, L. Yuan,
X. Hu and Y. Huang, Nano Energy, 2015, 12, 386–393.
328 Y. Tang, W. Zhang, L. Xue, X. Ding, T. Wang, X. Liu,
J. Liu, X. Li and Y. Huang, J. Mater. Chem. A, 2016, 4,
6036–6041.
329 S. Qiao, S. Dong, L. Yuan, T. Li, M. Ma, Y. Wu, Y. Hu, T. Qu
and S. Chong, J. Alloys Compd., 2023, 950, 169903.
330 A. Eftekhari, J. Power Sources, 2004, 126, 221–228.
331 X. Bie, K. Kubota, T. Hosaka, K. Chihara and S. Komaba,
J. Mater. Chem. A, 2017, 5,4325–4330.
332 L. Deng, J. Qu, X. Niu, J. Liu, J. Zhang, Y. Hong, M. Feng,
J. Wang, M. Hu, L. Zeng, Q. Zhang, L. Guo and Y. Zhu, Nat.
Commun., 2021, 12, 2167.
333 S. Chong, Y. Chen, Y. Zheng, Q. Tan, C. Shu, Y. Liu and
Z. Guo, J. Mater. Chem. A, 2017, 5, 22465–22471.
334 P. Padigi, J. Thiebes, M. Swan, G. Goncher, D. Evans and
R. Solanki, Electrochim. Acta, 2015, 166, 32–39.
335 Y.-h Zhu, Y.-b Yin, X. Yang, T. Sun, S. Wang, Y.-s Jiang,
J.-m Yan and X.-b Zhang, Angew. Chem., Int. Ed., 2017, 56,
7881–7885.
336 L. Xue, Y. Li, H. Gao, W. Zhou, X. Lu
¨, W. Kaveevivitchai,
A. Manthiram and J. B. Goodenough, J. Am. Chem. Soc.,
2017, 139, 2164–2167.
337 X. Jiang, T. Zhang, L. Yang, G. Li and J. Y. Lee, ChemElec-
troChem, 2017, 4, 2237–2242.
338 Z. Shadike, D.-R. Shi, W. Tian, M.-H. Cao, S.-F. Yang,
J. Chen and Z.-W. Fu, J. Mater. Chem. A, 2017, 5, 6393–6398.
339 X. Wu, Z. Jian, Z. Li and X. Ji, Electrochem. Commun., 2017,
77, 54–57.
340 G. He and L. F. Nazar, ACS Energy Lett., 2017, 2, 1122–1127.
341 X. Han, C. Chang, L. Yuan, T. Sun and J. Sun, Adv. Mater.,
2007, 19, 1616–1621.
342 A. Molina, N. Patil, E. Ventosa, M. Liras, J. Palma and
R. Marcilla, Adv. Funct. Mater., 2020, 30, 1908074.
343 Z. Ouyang, D. Tranca, Y. Zhao, Z. Chen, X. Fu, J. Zhu,
G. Zhai, C. Ke, E. Kymakis and X. Zhuang, ACS Appl. Mater.
Interfaces, 2021, 13, 9064–9073.
344 M. G. Mohamed, S. U. Sharma, C.-H. Yang, M. M. Samy,
A. A. K. Mohammed, S. V. Chaganti, J.-T. Lee and S. Wei-
Kuo, ACS Appl. Energy Mater., 2021, 4, 14628–14639.
345 Z. Chen, W. Li, Y. Dai, N. Xu, C. Su, J. Liu and C. Zhang,
Electrochim. Acta, 2018, 286, 187–194.
346 C. Zhang, X. Yang, W. Ren, Y. Wang, F. Su and J.-X. Jiang,
J. Power Sources, 2016, 317, 49–56.
Review Article Chem Soc Rev
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
This journal is © The Royal Society of Chemistry 2024 Chem. Soc. Rev.
347 K. S. Weeraratne, A. A. Alzharani and H. M. El-Kaderi, ACS
Appl. Mater. Interfaces, 2019, 11, 23520–23526.
348 K. Sakaushi, E. Hosono, G. Nickerl, T. Gemming, H. Zhou,
S. Kaskel and J. Eckert, Nat. Commun., 2013, 4, 1485.
349 K. Li, Y. Wang, B. Gao, X. Lv, Z. Si and H.-G. Wang,
J. Colloid Interface Sci., 2021, 601, 446–453.
350 B. Tian, J. Zheng, C. Zhao, C. Liu, C. Su, W. Tang,
X. Li and G.-H. Ning, J. Mater. Chem. A, 2019, 7,
9997–10003.
351 R. R. Kapaev, I. S. Zhidkov, E. Z. Kurmaev, K. J. Stevenson
and P. A. Troshin, J. Mater. Chem. A, 2019, 7, 22596–22603.
352 Z. Luo, L. Liu, J. Ning, K. Lei, Y. Lu, F. Li and J. Chen,
Angew. Chem., Int. Ed., 2018, 57, 9443–9446.
353 M. Wu, Y. Zhao, B. Sun, Z. Sun, C. Li, Y. Han, L. Xu, Z. Ge,
Y. Ren, M. Zhang, Q. Zhang, Y. Lu, W. Wang, Y. Ma and
Y. Chen, Nano Energy, 2020, 70, 104498.
354 K. Nakashima, T. Shimizu, Y. Kamakura, A. Hinokimoto,
Y. Kitagawa, H. Yoshikawa and D. Tanaka, Chem. Sci.,
2020, 11, 37–43.
355 R. Shi, L. Liu, Y. Lu, C. Wang, Y. Li, L. Li, Z. Yan and
J. Chen, Nat. Commun., 2020, 11, 178.
356 J. Duan, W. Wang, D. Zou, J. Liu, N. Li, J. Weng, L.-P. Xu,
Y. Guan, Y. Zhang and P. Zhou, ACS Appl. Mater. Interfaces,
2022, 14, 31234–31244.
357 J. Lee, H. Lim, J. Park, M.-S. Kim, J.-W. Jung, J. Kim and
I.-D. Kim, Adv. Energy Mater, 2023, 13, 2300442.
358 G. Fe
´rey, F. Millange, M. Morcrette, C. Serre, M.-L. Doublet,
J.-M. Grene
`che and J.-M. Tarascon, Angew. Chem., Int. Ed.,
2007, 46, 3259–3263.
359 T. L. A. Nguyen, R. Demir-Cakan, T. Devic, M. Morcrette,
T. Ahnfeldt, P. Auban-Senzier, N. Stock, A.-M. Goncalves,
Y. Filinchuk, J.-M. Tarascon and G. Fe
´rey, Inorg. Chem.,
2010, 49, 7135–7143.
360 B. Tian, G.-H. Ning, Q. Gao, L.-M. Tan, W. Tang, Z. Chen,
C. Su and K. P. Loh, ACS Appl. Mater. Interfaces, 2016, 8,
31067–31075.
361 S. Zheng, Y. Chen, Y. Xu, F. Yi, Y. Zhu, Y. Liu, J. Yang and
C. Wang, ACS Nano, 2013, 7, 10995–11003.
362 Z. Li, X. Li, Y. Liao, X. Li and W. Li, J. Power Sources,2016,
334, 23–30.
363 Y. Liu, L. Si, X. Zhou, X. Liu, Y. Xu, J. Bao and Z. Dai,
J. Mater. Chem. A, 2014, 2, 17735–17739.
364 Y. Yan, M. Shi, Y. Wei, C. Zhao, M. Carnie, R. Yang and
Y. Xu, J. Alloys Compd., 2018, 738, 16–24.
365 Z. Li, L. Yuan, Z. Yi, Y. Sun, Y. Liu, Y. Jiang, Y. Shen, Y. Xin,
Z. Zhang and Y. Huang, Adv. Energy Mater., 2014,
4, 1301473.
366 M. H. Aboonasr Shiraz, H. Zhu, Y. Liu, X. Sun and J. Liu,
J. Power Sources, 2019, 438, 227059.
367 Q. Guo, S. Li, X. Liu, H. Lu, X. Chang, H. Zhang, X. Zhu,
Q. Xia, C. Yan and H. Xia, Adv. Sci., 2020, 7, 1903246.
368 D. Zhao, S. Jiang, S. Yu, J. Ren, Z. Zhang, S. Liu,
X. Liu, Z. Wang, Y. Wu and Y. Zhang, Carbon, 2023, 201,
864–870.
369 Q. Guo, S. Sun, K.-I. Kim, H. Zhang, X. Liu, C. Yan and
H. Xia, Carbon Energy, 2021, 3, 440–448.
370 D. Zhao, S. Ge-Zhang, Z. Zhang, H. Tang, Y. Xu, F. Gao,
X. Xu, S. Liu, J. Zhou, Z. Wang, Y. Wu, X. Liu and Y. Zhang,
ACS Appl. Mater. Interfaces, 2022, 14, 54662–54669.
371 P. Xiong, X. Han, X. Zhao, P. Bai, Y. Liu, J. Sun and Y. Xu,
ACS Nano, 2019, 13, 2536–2543.
372 X. Han, M. Wang, J. Yu and S. Wang, Mater. Adv., 2022, 3,
1652–1659.
373 Y. Xiao, Y. Miao, S. Wan, Y.-K. Sun and S. Chen, Small,
2022, 18, 2202582.
374 H.-S. Zhao, Y.-L. Qi, K. Liang, W.-K. Zhu, H.-B. Wu, J.-B. Li
and Y.-R. Ren, Rare Met., 2022, 41, 1284–1293.
375 M. Nazarian-Samani, M. Nazarian-Samani, S. Haghighat-
Shishavan and K.-B. Kim, Energy Storage Mater., 2021, 36,
229–241.
376 Z. Wang, H. Shi, S. Yang, Z. Cai, H. Lu, L. Jia, M. Hu, H. He
and K. Zhou, J. Alloys Compd., 2021, 888, 161553.
377 Y. Zhang, M. Zhang, Y. Liu, H. Zhu, L. Wang, Y. Liu,
M. Xue, B. Li and X. Tao, Electrochim. Acta, 2020,
330, 135299.
378 C. M. Park and H. J. Sohn, Adv. Mater., 2007, 19,
2465–2468.
379 C. Xia, Y. Zhou, D. B. Velusamy, A. A. Farah, P. Li, Q. Jiang,
I. N. Odeh, Z. Wang, X. Zhang and H. N. Alshareef, Nano
Lett.,2018, 18, 1506–1515.
380 U. Chothe, C. Ugale, M. Kulkarni and B. Kale, Crystals,
2021, 11, 660.
381 Z. Hu, Q. Liu, W. Sun, W. Li, Z. Tao, S.-L. Chou, J.
Chen and S.-X. Dou, Inorganic Chemistry Frontiers, 2016,
3, 532–535.
382 R. Zhou, H. Wang, J. Chang, C. Yu, H. Dai, Q. Chen,
J. Zhou, H. Yu, G. Sun and W. Huang, ACS Appl. Mater.
Interfaces, 2021, 13, 17459–17466.
383 F. Wang, Y. Li, X. Liang, Y. Liu, Q. Chen and M. Chen,
Mater. Chem. Phys., 2022, 278, 125681.
384 P. Lian, Y. Dong, Z.-S. Wu, S. Zheng, X. Wang, W. Sen,
C. Sun, J. Qin, X. Shi and X. Bao, Nano Energy, 2017, 40,
1–8.
385 P. A. Maughan, L. Bouscarrat, V. R. Seymour, S. Shao,
S. J. Haigh, R. Dawson, N. Tapia-Ruiz and N. Bimbo,
Nanoscale Adv., 2021, 3, 3145–3158.
386 S. Kajiyama, L. Szabova, K. Sodeyama, H. Iinuma,
R. Morita, K. Gotoh, Y. Tateyama, M. Okubo and
A. Yamada, ACS Nano, 2016, 10, 3334–3341.
387 Y. Xie, Y. Dall’Agnese, M. Naguib, Y. Gogotsi, M. W.
Barsoum, H. L. Zhuang and P. R. C. Kent, ACS Nano,
2014, 8, 9606–9615.
388 X. Wang, S. Kajiyama, H. Iinuma, E. Hosono, S. Oro,
I. Moriguchi, M. Okubo and A. Yamada, Nat. Commun.,
2015, 6, 6544.
389 Y. Dall’Agnese, P.-L. Taberna, Y. Gogotsi and P. Simon,
J. Phys. Chem. Lett., 2015, 6, 2305–2309.
390 S. Imtiaz, J. Zhang, Z. A. Zafar, S. N. Ji, T. Z. Huang,
J. A. Anderson, Z. L. Zhang and Y. H. Huang, Science
China-Materials, 2016, 59, 389–407.
391 Z. Lu, C.-T. Yip, L. Wang, H. Huang and L. Zhou, Chem-
PlusChem, 2012, 77, 991–1000.
Chem Soc Rev Review Article
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2024
392 K. Su, T. Jin, C. H. Zhang, R. Wang, S. Yuan, N. W. Li and
L. Yu, J. Chem. Eng., 2022, 450, 138049.
393 J. Han, I. Johnson, Z. Lu, A. Kudo and M. Chen, Nano Lett.,
2021, 21, 6504–6510.
394 L. Zhong, W. Zhang, S. Sun, L. Zhao, W. Jian, X. He,
Z. Xing, Z. Shi, Y. Chen, H. N. Alshareef and X. Qiu, Adv.
Funct. Mater., 2023, 33, 2211872.
395 J. Wu, X. Zhang, Z. Li, C. Yang, W. Zhong, W. Li, C. Zhang,
N. Yang, Q. Zhang and X. Li, Adv. Funct. Mater., 2020,
30, 2004348.
396 N. Xiao, X. Zhang, C. Liu, Y. Wang, H. Li and J. Qiu,
Carbon, 2019, 147, 574–581.
397 D. Yoon, J. Hwang, W. Chang and J. Kim, ACS Appl. Mater.
Interfaces, 2018, 10, 569–581.
398 J. Yang, X. Wang, W. Dai, X. Lian, X. Cui, W. Zhang,
K. Zhang, M. Lin, R. Zou, K. P. Loh, Q.-H. Yang and
W. Chen, Nano-Micro Lett., 2021, 13, 98.
399 S. Zhang, W. Huang, P. Hu, C. Huang, C. Shang, C.
Zhang, R. Yang and G. Cui, J. Mater. Chem. A, 2015, 3,
1896–1901.
400 T. Yang, C. Zhang, W. Ma, X. Gao, C. Yan, F. Wang and
J.-X. Jiang, Solid State Ion., 2020, 347, 115247.
401 C. Zhang, Y. Qiao, P. Xiong, W. Ma, P. Bai, X. Wang, Q. Li,
J. Zhao, Y. Xu, Y. Chen, J. H. Zeng, F. Wang, Y. Xu and
J.-X. Jiang, ACS Nano, 2019, 13, 745–754.
402 S.-Y. Li, W.-H. Li, X.-L. Wu, Y. Tian, J. Yue and G. Zhu,
Chem. Sci., 2019, 10, 7695–7701.
403 Y. Han, P. Qi, X. Feng, S. Li, X. Fu, H. Li, Y. Chen, J. Zhou,
X. Li and B. Wang, ACS Appl. Mater. Interfaces, 2015, 7,
2178–2182.
404 C. Yan, H. Zhao, J. Li, H. Jin, L. Liu, W. Wu, J. Wang, Y. Lei
and S. Wang, Small, 2020, 16, 1907141.
405 W. Li, S. Hu, X. Luo, Z. Li, X. Sun, M. Li, F. Liu and Y. Yu,
Adv. Mater., 2017, 29, 1605820.
406 S. Yao, J. Cui, J. Huang, J.-Q. Huang, W. G. Chong, L. Qin,
Y.-W. Mai and J.-K. Kim, Adv. Energy Mater, 2018,
8, 1702267.
407 S.-A. He, Q. Liu, Z. Cui, K. Xu, R. Zou, W. Luo and M. Zhu,
Small, 2022, 18, 2105866.
408 J. Xiao, Z. Cai, T. Muhmood, X. Hu, S. Lin and X. Hu, Small,
2022, 18, 2106930.
409 Y. Liu, S. Sun, Y. Wang, C. Wang, Q.-C. Wang, J. Han and
R. Guo, J. Power Sources, 2022, 549, 232087.
410 L. Wu, L. Wang, X. Cheng, M. Ma, Y. Wu, X. Wu,
H. Yang, Y. Yu and C. He, Nano Res., 2022, 15,
2147–2156.
411 Z. Yan, Z. Sun, A. Li, H. Liu, Z. Guo, L. Zhao, J. Feng and
L. Qian, J. Chem. Eng., 2022, 429, 132249.
412 W. Li, H. Li, Z. Lu, L. Gan, L. Ke, T. Zhai and H. Zhou,
Energy Environ. Sci., 2015, 8, 3629–3636.
413 L. Z. Wenchao Bi, Jian Chen, Ruixue Tian, Hao Huang and
Man Yao, Acta Chim, 2022, 80, 756–764.
414 F. Song, R. Zhang, X. Zhang, J. Qin and R. Liu, Energy Fuels,
2020, 34, 13032–13037.
415 Y.-N. Zhang, Y. Zhou, J. Su, Y.-F. Long, X.-Y. Lv, H.-X. Kuai
and Y.-X. Wen, Appl. Surf. Sci., 2022, 585, 152763.
416 Y. Zhao, T. Sun, Q. Yin, J. Zhang, S. Zhang, J. Luo, H. Yan,
L. Zheng, J. Han and M. Wei, J. Mater. Chem. A, 2019, 7,
15371–15377.
417 P. Lian, X. Zhu, S. Liang, Z. Li, W. Yang and H. Wang,
Electrochim. Acta, 2010, 55, 3909–3914.
418 Y. Sun, J. Tang, K. Zhang, J. Yuan, J. Li, D.-M. Zhu,
K. Ozawa and L.-C. Qin, Nanoscale, 2017, 9, 2585–2595.
419 G. Ma, K. Huang, J.-S. Ma, Z. Ju, Z. Xing and Q.-C. Zhuang,
J. Mater. Chem. A, 2017, 5, 7854–7861.
420 Z. Chen, Y. Liu, Y. Zhang, F. Shen, G. Yang, L. Wang,
X. Zhang, Y. He, L. Luo and S. Deng, Mater. Lett., 2018, 229,
134–137.
421 Z.-L. Xu, G. Yoon, K.-Y. Park, H. Park, O. Tamwattana,
S. Joo Kim, W. M. Seong and K. Kang, Nat. Commun., 2019,
10, 2598.
422 Y. Wen, K. He, Y. Zhu, F. Han, Y. Xu, I. Matsuda,
Y. Ishii, J. Cumings and C. Wang, Nat. Commun., 2014,
5, 4033.
423 M. G. Kim, S. Lee and J. Cho, J. Electrochem. Soc., 2009,
156, A89.
424 F. Han, C. Zhang, J. Yang, G. Ma, K. He and X. Li, J. Mater.
Chem. A, 2016, 4, 12781–12789.
425 Y. Kim, H. Hwang, C. S. Yoon, M. G. Kim and J. Cho, Adv.
Mater., 2007, 19, 92–96.
426 Q. Zhang, P. Man, B. He, C. Li, Q. Li, Z. Pan, Z. Wang,
J. Yang, Z. Wang, Z. Zhou, X. Lu, Z. Niu, Y. Yao and L. Wei,
Nano Energy, 2020, 67, 104212.
427 P. Senguttuvan, G. Rousse, V. Seznec, J.-M. Tarascon and
M. R. Palacı
´n, Chem. Mater., 2011, 23, 4109–4111.
428 T. Yuan, Z. Tan, C. Ma, J. Yang, Z.-F. Ma and S. Zheng, Adv.
Energy Mater., 2017, 7, 1601625.
429 E. Griffin, L. Mogg, G.-P. Hao, G. Kalon, C. Bacaksiz,
G. Lopez-Polin, T. Zhou, V. Guarochico, J. Cai and
C. Neumann, ACS Nano, 2020, 14, 7280–7286.
430 J. Chmiola, C. Largeot, P. L. Taberna, P. Simon and
Y. Gogotsi, Angew. Chem., Int. Ed., 2008, 47, 3392–3395.
431 C. Lu, C. Hu, C. L. Ritt, X. Hua, J. Sun, H. Xia, Y. Liu,
D.-W. Li, B. Ma and M. Elimelech, J. Am. Chem. Soc., 2021,
143, 14242–14252.
432 Y. Li, X. Jin, X. Yan, X. Ai, X. Yang, Z.-J. Zheng, K. Huang,
G. Zhao, Y. Yang, M. Wu and K.-G. Zhou, Nano Res, 2023,
16, 10913–10921.
433 H. X. Yang, Z. K. Liu, Y. Wang, N. W. Li and L. Yu, Adv.
Funct. Mater., 2023, 33, 2209837.
434 X. Li, H. Zhang, P. Wang, J. Hou, J. Lu, C. D. Easton,
X. Zhang, M. R. Hill, A. W. Thornton and J. Z. Liu, Nat.
Commun., 2019, 10, 2490.
435 Z. Chang, H. Yang, X. Zhu, P. He and H. Zhou, Nat.
Commun., 2022, 13, 1510.
436 L. Du, B. Zhang, W. Deng, Y. Cheng, L. Xu and L. Mai, Adv.
Energy Mater., 2022, 12, 2200501.
437 W. He, D. Li, S. Guo, Y. Xiao, W. Gong, Q. Zeng, Y. Ouyang,
X. Li, H. Deng and C. Tan, Energy Storage Mater., 2022, 47,
271–278.
438 G. S. MacGlashan, Y. G. Andreev and P. G. Bruce, Nature,
1999, 398, 792–794.
Review Article Chem Soc Rev
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
This journal is © The Royal Society of Chemistry 2024 Chem. Soc. Rev.
439 C. Yang, Q. Wu, W. Xie, X. Zhang, A. Brozena, J. Zheng,
M. N. Garaga, B. H. Ko, Y. Mao and S. He, Nature, 2021,
598, 590–596.
440 Q. Zeng, J. Wang, X. Li, Y. Ouyang, W. He, D. Li, S. Guo,
Y. Xiao, H. Deng, W. Gong, Q. Zhang and S. Huang, ACS
Energy Lett., 2021, 6, 2434–2441.
441 A. Hu, Z. Sun, Q. Hou, J. Duan, C. Li, W. Dou, J. Fan,
M. Zheng and Q. Dong, Small, 2022, 18, 2205571.
442 Y. Kato, S. Hori, T. Saito, K. Suzuki, M. Hirayama,
A. Mitsui, M. Yonemura, H. Iba and R. Kanno, Nat. Energy.,
2016, 1, 1–7.
443 J. A. Dawson, P. Canepa, T. Famprikis, C. Masquelier and
M. S. Islam, J. Am. Chem. Soc., 2018, 140, 362–368.
444 G. Redhammer, D. Rettenwander, S. Pristat, E. Dashjav,
C. Kumar, D. Topa and F. Tietz, Solid State Sci., 2016, 60,
99–107.
445 G. Yang, C. Chanthad, H. Oh, I. A. Ayhan and Q. Wang,
J. Mater. Chem. A, 2017, 5, 18012–18019.
446 M. F. Majid, H. F. Mohd Zaid, C. F. Kait, A. Ahmad and
K. Jumbri, Nanomaterials, 2022, 12(7), 1076.
447 T. Zhou, Y. Zhao, J. W. Choi and A. Coskun, Angew. Chem.,
Int. Ed., 2021, 60, 22791–22796.
448 D.-J. Yoo, K. J. Kim and J. W. Choi, Adv. Energy Mater.,
2018, 8, 1702744.
449 W. Lei, C. Zhang, R. Qiao, M. Ravivarma, H. Chen,
F. B. Ajdari, M. Salavati-Niasari and J. Song, ACS Appl.
Energy Mater., 2023, 6, 4363–4371.
450 Z. Wang, H. Zhou, C. Meng, L. Zhang, Y. Cai and A. Yuan,
ChemElectroChem, 2020, 7, 2660–2664.
451 E. M. Miner, S. S. Park and M. Dinca
˘,J. Am. Chem. Soc.,
2019, 141, 4422–4427.
452 T. Wang, X. Zhang, N. Yuan and C. Sun, J. Chem. Eng.,
2023, 451, 138819.
453 B. M. Wiers, M.-L. Foo, N. P. Balsara and J. R. Long, J. Am.
Chem. Soc., 2011, 133, 14522–14525.
454 T. Zhao, W. Kou, Y. Zhang, W. Wu, W. Li and J. Wang,
J. Power Sources, 2023, 554, 232349.
455 L. Liu and C. Sun, ChemElectroChem, 2020, 7,707–715.
456 K. Fujie, R. Ikeda, K. Otsubo, T. Yamada and H. Kitagawa,
Chem. Mater., 2015, 27, 7355–7361.
457 Z. Wang, R. Tan, H. Wang, L. Yang, J. Hu, H. Chen and
F. Pan, Adv. Mater., 2018, 30, 1704436.
458 K. Wang, L. Yang, Z. Wang, Y. Zhao, Z. Wang, L. Han,
Y. Song and F. Pan, Chem. comm., 2018, 54, 13060–13063.
459 Z. Liu, Z. Hu, X. Jiang, X. Wang, Z. Li, Z. Chen, Y. Zhang
and S. Zhang, Small, 2022, 18, 2203011.
460 Z. Wang, Z. Wang, L. Yang, H. Wang, Y. Song, L. Han,
K. Yang, J. Hu, H. Chen and F. Pan, Nano Energy, 2018, 49,
580–587.
461 R. Zettl and I. Hanzu, Front. energy res., 2021, 9, 714698.
462 T. Jiang, P. He, G. Wang, Y. Shen, C.-W. Nan and L.-Z. Fan,
Adv. Energy Mater., 2020, 10, 1903376.
463 W. Lu, M. Xue and C. Zhang, Energy Storage Mater., 2021,
39, 108–129.
464 S. Liu, W. Liu, D. Ba, Y. Zhao, Y. Ye, Y. Li and J. Liu, Adv.
Mater., 2023, 35, 2110423.
465 W. Liu, D. Lin, J. Sun, G. Zhou and Y. Cui, ACS Nano, 2016,
10, 11407–11413.
466 Y. Zhu, J. Cao, H. Chen, Q. Yu and B. Li, J. Mater. Chem. A,
2019, 7, 6832–6839.
467 K. He, C. Chen, R. Fan, C. Liu, C. Liao, Y. Xu, J. Tang and
R. K. Y. Li, Compos. Sci. Technol., 2019, 175, 28–34.
468 K. West, B. Zachau-Christiansen, T. Jacobsen, E. Hiort-
Lorenzen and S. Skaarup, Br. Polym. J., 1988, 20, 243–246.
469 J. Serra Moreno, M. Armand, M. B. Berman, S. G.
Greenbaum, B. Scrosati and S. Panero, J. Power Sources,
2014, 248, 695–702.
470 X. Qi, Q. Ma, L. Liu, Y.-S. Hu, H. Li, Z. Zhou, X. Huang and
L. Chen, ChemElectroChem, 2016, 3, 1741–1745.
471 A. Bitner-Michalska, G. M. Nolis, G. Z
˙ukowska, A.
Zalewska, M. Poterała, T. Trzeciak, M. Dranka, M. Kalita,
P. Jankowski, L. Niedzicki, J. Zachara, M. Marcinek and
W. Wieczorek, Sci. Rep., 2017, 7, 40036.
472 C. V. Subba Reddy, A. P. Jin, Q. Y. Zhu, L. Q. Mai and
W. Chen, Eur. Phys. J. E, 2006, 19, 471–476.
473 P. B. Bhargav, V. M. Mohan, A. K. Sharma and
V. V. R. N. Rao, J. Appl. Polym. Sci., 2008, 108, 510–517.
474 H. Gao, S. Xin, L. Xue and J. B. Goodenough, Chem, 2018,
4, 833–844.
475 S. Song, H. M. Duong, A. M. Korsunsky, N. Hu and L. Lu,
Sci. Rep, 2016, 6, 32330.
476 Q. Ma, M. Guin, S. Naqash, C.-L. Tsai, F. Tietz and
O. Guillon, Chem. Mater., 2016, 28, 4821–4828.
477 X. Chi, Y. Liang, F. Hao, Y. Zhang, J. Whiteley, H. Dong,
P. Hu, S. Lee and Y. Yao, Angew. Chem., Int. Ed., 2018, 57,
2630–2634.
478 N. Tanibata, K. Noi, A. Hayashi and M. Tatsumisago, RSC
Adv., 2014, 4, 17120–17123.
479 Z. Yu, S.-L. Shang, J.-H. Seo, D. Wang, X. Luo, Q. Huang,
S. Chen, J. Lu, X. Li, Z.-K. Liu and D. Wang, Adv. Mater.,
2017, 29, 1605561.
480 I.-H. Chu, C. S. Kompella, H. Nguyen, Z. Zhu, S. Hy,
Z. Deng, Y. S. Meng and S. P. Ong, Sci. Rep., 2016, 6, 33733.
481 L. Zhang, K. Yang, J. Mi, L. Lu, L. Zhao, L. Wang, Y. Li and
H. Zeng, Adv. Energy Mater., 2015, 5, 1501294.
482 T. J. Udovic, M. Matsuo, W. S. Tang, H. Wu, V. Stavila,
A. V. Soloninin, R. V. Skoryunov, O. A. Babanova,
A. V. Skripov, J. J. Rush, A. Unemoto, H. Takamura and
S.-I. Orimo, Adv. Mater, 2014, 26, 7622–7626.
483 L. Duche
ˆne, R. S. Ku
¨hnel, D. Rentsch, A. Remhof,
H. Hagemann and C. Battaglia, Chem. comm., 2017, 53,
4195–4198.
484 Z. Zhang, Q. Zhang, C. Ren, F. Luo, Q. Ma, Y.-S. Hu,
Z. Zhou, H. Li, X. Huang and L. Chen, J. Mater. Chem. A,
2016, 4, 15823–15828.
485 S. Li, M. Jiang, Y. Xie, H. Xu, J. Jia and J. Li, Adv. Mater.,
2018, 30, e1706375.
486 A. Chandra, Ind. J. Phys., 2016, 90, 759–765.
487 P. Kesharwani, D. K. Sahu, M. Sahu, T. B. Sahu and
R. C. Agrawal, Ionics, 2017, 23, 2823–2827.
488 J. Siva Kumar, M. Jaipal Reddy and U. V. Subba Rao,
J. Mater. Sci., 2006, 41,6171–6173.
Chem Soc Rev Review Article
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2024
489 C. Zhao, L. Liu, X. Qi, Y. Lu, F. Wu, J. Zhao, Y. Yu, Y.-S. Hu
and L. Chen, Adv. Energy Mater, 2018, 8, 1703012.
490 P. Kesharwani, D. K. Sahu, Y. K. Mahipal and
R. C. Agrawal, Mater. Chem. Phys., 2017, 193, 524–531.
491 H. Yuan, H. Li, T. Zhang, G. Li, T. He, F. Du and S. Feng,
J. Mater. Chem. A, 2018, 6, 8413–8418.
492 Z. Li, L. Wang, Y. Liu, M. Yu, B. Liu, Y. Men, Z. Sun, W. Hu
and G. Zhu, Small, 2023, 2302818.
493 K. Shen, Z. Wang, X. Bi, Y. Ying, D. Zhang, C. Jin, G. Hou,
H. Cao, L. Wu, G. Zheng, Y. Tang, X. Tao and J. Lu, Adv.
Energy Mater, 2019, 9, 1900260.
494 Z. Lu, H. Yang, Y. Guo, P. He, S. Wu, Q.-H. Yang and
H. Zhou, Angew. Chem., Int. Ed., 2022, 61, e202206340.
495 R. Tan, A. Wang, R. Malpass-Evans, R. Williams, E. W.
Zhao,T.Liu,C.Ye,X.Zhou,B.P.Darwich,Z.Fan,L.Turcani,
E.Jackson,L.Chen,S.Y.Chong,T.Li,K.E.Jelfs,A.I.Cooper,
N.P.Brandon,C.P.Grey,N.B.McKeownandQ.Song,Nat.
Mater., 2020, 19, 195–202.
496 J. C. Barbosa, D. M. Correia, A. Fidalgo-Marijuan, R.
Gonçalves, S. Ferdov, V. de Zea Bermudez, S. Lanceros-
Mendez and C. M. Costa, ACS Appl. Mater. Interfaces, 2023,
15, 32301–32312.
497 H. Jamal, F. Khan, S. Hyun, S. W. Min and J. H. Kim,
J. Mater. Chem. A, 2021, 9, 4126–4137.
498 H. Jamal, F. Khan, H.-R. Si and J. H. Kim, J. Mater. Chem. A,
2021, 9, 27304–27319.
499 Y.Feng,Y.Li,J.Lin,H.Wu,L.Zhu,X.Zhang,L.Zhang,
C.-F.Sun,M.WuandY.Wang,Nat. Commun., 2023, 14, 3639.
500 T. Kim, S. Hyeok Ahn, Y.-Y. Song, B. Jin Park, C. Lee,
A. Choi, M.-H. Kim, D.-H. Seo, S.-K. Jung and H.-W. Lee,
Angew. Chem., Int. Ed., 2023, e202309852.
501 A. Hayashi, K. Noi, A. Sakuda and M. Tatsumisago, Nat.
Commun., 2012, 3, 856.
502 R. Schlem, A. Banik, M. Eckardt, M. Zobel and W. G. Zeier,
ACS Appl. Energy Mater., 2020, 3, 10164–10173.
503 T. Fuchs, S. P. Culver, P. Till and W. G. Zeier, ACS Energy
Lett., 2020, 5, 146–151.
504 E. A. Wu, S. Banerjee, H. Tang, P. M. Richardson,
J.-M. Doux, J. Qi, Z. Zhu, A. Grenier, Y. Li, E. Zhao,
G. Deysher, E. Sebti, H. Nguyen, R. Stephens, G. Verbist,
K. W. Chapman, R. J. Cle
´ment, A. Banerjee, Y. S. Meng and
S. P. Ong, Nat. Commun., 2021, 12, 1256.
505 J. Shao, H. Ao, L. Qin, J. Elgin, C. E. Moore, Y. Khalifa,
S. Zhang and Y. Wu, Adv. Mater., 2023, 2306809.
506 D. Sarkar, A. Bhattacharya, J. Meyer, A. M. Kirchberger,
V. Mishra, T. Nilges and V. K. Michaelis, J. Am. Chem. Soc.,
2023, 145, 19727–19745.
507 Y. Li, Z. Deng, J. Peng, J. Gu, E. Chen, Y. Yu, J. Wu, X. Li,
J. Luo, Y. Huang, Y. Xu, Z. Gao, C. Fang, J. Zhu, Q. Li,
J. Han and Y. Huang, ACS Appl. Mater. Interfaces, 2018, 10,
15760–15766.
508 Y. Cheng, X. Liu, Y. Guo, G. Dong, X. Hu, H. Zhang, X. Xiao,
Q. Liu, L. Xu and L. Mai, Adv. Mater., 2023, 2303226.
509 J.-H. Lee, H. Lee, J. Lee, T. W. Kang, J. H. Park, J.-H. Shin,
H. Lee, D. Majhi, S. U. Lee and J.-H. Kim, ACS Nano, 2023,
17, 17372–17382.
510 L. Du, B. Zhang, C. Yang, L. Cui, L. Mai and L. Xu, Energy
Storage Mater., 2023, 61, 102914.
511 R. Xu, Y. Kang, W. Zhang, B. Pan and X. Zhang, Nat.
Commun., 2023, 14, 4907.
512 X. Wang, Y. Luo, H. Wang, C. Wu, Z. Zhang and J. Li,
J. Electroanal. Chem., 2021, 897, 115564.
513 H. Sun, G. Zhu, Y. Zhu, M. C. Lin, H. Chen, Y. Y. Li,
W. H. Hung, B. Zhou, X. Wang, Y. Bai, M. Gu, C. L. Huang,
H. C. Tai, X. Xu, M. Angell, J. J. Shyue and H. Dai, Adv.
Mater., 2020, e2001741, DOI: 10.1002/adma.202001741.
514 Y. Zang, F. Pei, J. Huang, Z. Fu, G. Xu and X. Fang, Adv.
Energy Mater., 2018, 8, 1802052.
515 Z. Jin, K. Xie and X. Hong, RSC Adv., 2013, 3, 8889–8898.
516 Y. Suharto, Y. Lee, J.-S. Yu, W. Choi and K. J. Kim, J. Power
Sources, 2018, 376, 184–190.
517 S. Bai, X. Liu, K. Zhu, S. Wu and H. Zhou, Nat. Energy.,
2016, 1, 16094.
518 H. Kwon, H.-J. Choi, J.-K. Jang, J. Lee, J. Jung, W. Lee,
Y. Roh, J. Baek, D. J. Shin, J.-H. Lee, N.-S. Choi, Y. S. Meng
and H.-T. Kim, Nat. Commun., 2023, 14, 4047.
519 C. Yin, Z. Li, D. Zhao, J. Yang, Y. Zhang, Y. Du and
Y. Wang, ACS Nano, 2022, 16, 14178–14187.
520 H. Zhang, C. Lin, X. Hu, B. Zhu and D. Yu, ACS Appl. Mater.
Interfaces, 2018, 10, 12708–12715.
521 C. Ye, A. Wang, C. Breakwell, R. Tan, C. Grazia Bezzu, E.
Hunter-Sellars, D. R. Williams, N. P. Brandon, P. A. A.
Klusener, A. R. Kucernak, K. E. Jelfs, N. B. McKeown and
Q. Song, Nat. Commun., 2022, 13, 3184.
522 C. Qi, L. Xu, J. Wang, H. Li, C. Zhao, L. Wang and T. Liu,
ACS Sustainable Chem. Eng., 2020, 8, 12968–12975.
523 A. Wang, R. Tan, D. Liu, J. Lu, X. Wei, A. Alvarez-Fernandez,
C.Ye,C.Breakwell,S.Guldin,A.R.Kucernak,K.E.Jelfs,
N. P. Brandon, N. B. McKeown and Q. Song, Adv. Mater.,
2023, 35, 2210098.
524 Y. Huai, J. Gao, Z. Deng and J. Suo, Ionics, 2010, 16,
603–611.
525 J. H. Lee, J. Manuel, H. Choi, W. H. Park and J.-H. Ahn,
Polymer, 2015, 68, 335–343.
526 S. Janakiraman, A. Surendran, S. Ghosh, S. Anandhan and
A. Venimadhav, Solid State Ion., 2016, 292, 130–135.
527 S. Janakiraman, M. Khalifa, R. Biswal, S. Ghosh,
S. Anandhan and A. Venimadhav, J. Power Sources, 2020,
460, 228060.
528 H. Fei, Y. Liu, Y. An, X. Xu, J. Zhang, B. Xi, S. Xiong and
J. Feng, J. Power Sources, 2019, 433, 226697.
529 H. Fei, Y. Liu, Y. An, X. Xu, G. Zeng, Y. Tian, L. Ci, B. Xi,
S. Xiong and J. Feng, J. Power Sources, 2018, 399, 294–298.
530 T. C. Mendes, X. Zhang, Y. Wu, P. C. Howlett, M. Forsyth
and D. R. Macfarlane, ACS Sustain. Chem. Eng., 2019, 7,
3722–3726.
531 S. Chen, J. Zheng, L. Yu, X. Ren, M. H. Engelhard, C. Niu,
H. Lee, W. Xu, J. Xiao, J. Liu and J.-G. Zhang, Joule, 2018, 2,
1548–1558.
532 H. Yang, J. Bright, B. Chen, P. Zheng, X. Gao, B. Liu,
S. Kasani, X. Zhang and N. Wu, J. Mater. Chem. A, 2020,
8, 7261–7272.
Review Article Chem Soc Rev
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
This journal is © The Royal Society of Chemistry 2024 Chem. Soc. Rev.
533 X. Chen, Y. Xie, Y. Ling, J. Zhao, Y. Xu, Y. Tong, S. Li and
Y. Wang, Mater. Des., 2020, 192, 108760.
534 B. Zhou, Y. H. Jo, R. Wang, D. He, X. Zhou, X. Xie and
Z. Xue, J. Mater. Chem. A, 2019, 7, 10354–10362.
535 Z. Li, T. Chen and Y. Liao, Ionics, 2015, 21, 2763–2770.
536 R. Subramani, Y.-H. Tseng, Y.-L. Lee, C.-C. Chiu, S.-S. Hou
and H. Teng, J. Mater. Chem. A, 2019, 7, 12244–12252.
537 X. Ren, L. Zou, X. Cao, M. H. Engelhard, W. Liu, S. D.
Burton, H. Lee, C. Niu, B. E. Matthews, Z. Zhu, C. Wang,
B. W. Arey, J. Xiao, J. Liu, J.-G. Zhang and W. Xu, Joule,
2019, 3, 1662–1676.
538 H. Li, Y.-M. Chen, X.-T. Ma, J.-L. Shi, B.-K. Zhu and
L.-P. Zhu, J. Membr. Sci., 2011, 379, 397–402.
539 H.-L. Guo, H. Sun, Z.-L. Jiang, C.-S. Luo, M.-Y. Gao,
M.-H. Wei, J.-Y. Hu, W.-K. Shi, J.-Y. Cheng and H.-J.
Zhou, J. Mater. Sci., 2019, 54, 4874–4883.
540 G. Fu, M. D. Soucek and T. Kyu, Solid State Ion., 2018, 320,
310–315.
541 P. Shi, T. Li, R. Zhang, X. Shen, X. B. Cheng, R. Xu,
J. Q. Huang, X. R. Chen, H. Liu and Q. Zhang, Adv. Mater.,
2019, 31, e1807131.
542 P. Chen, J. Shen, T. Wang, M. Dai, C. Si, J. Xie, M. Li,
X. Cong and X. Sun, J. Power Sources, 2018, 400, 325–332.
543 T. Song, C. Wang and C.-S. Lee, Carbon Neutraliz., 2022, 1,
68–92.
544 K. Mizushima, P. C. Jones, P. J. Wiseman and J. B.
Goodenough, Mater. Res. Bull., 1980, 15, 783–789.
545 I. Bezza, E. Luais, F. Ghamouss, M. Zaghrioui, F. Tran-van
and J. Sakai, J. Alloys Compd., 2019, 805, 19–25.
546 H. Arai, S. Okada, Y. Sakurai and J.-I. Yamaki, Solid State
Ion., 1997, 95, 275–282.
547 S.-H. Wu and M.-T. Yu, J. Power Sources, 2007, 165,
660–665.
548 Y. Sakurai, H. Arai, S. Okada and J.-I. Yamaki, J. Power
Sources, 1997, 68, 711–715.
549 J. Molenda, C. Delmas and P. Hagenmuller, Solid State
Ion., 1983, 9–10,431–435.
550 A. Eftekhari, Z. Jian and X. Ji, ACS Appl. Mater. Interfaces,
2017, 9, 4404–4419.
551 J. Li, H. Hu, J. Wang and Y. Xiao, Carbon Neutraliz, 2022, 1,
96–116.
552 J. Su, Y. Pei, Z. Yang and X. Wang, RSC Adv., 2015, 5,
27229–27234.
553 B. Hu, F. Geng, C. Zhao, B. Doumert, J. Tre
´bosc, O. Lafon,
C. Li, M. Shen and B. Hu, ACS Appl. Mater. Interfaces, 2020,
12, 41485–41494.
554 S. Yun, S. Muhammad, J. Choi, W. Lee, J. Yu, H. Lee,
Y. Lee and W.-S. Yoon, ACS Energy Lett, 2022, 7, 3989–3996.
555 X. Zhang, X. Yang, G. Sun, S. Yao, Y. Xie, W. Zhang, C. Liu,
X. Wang, R. Yang, X. Jin, Z. X. Shen, H. J. Fan and F. Du,
Adv. Funct. Mater., 2022, 32, 2204318.
556 Z. Lu and J. R. Dahn, J. Electrochem. Soc., 2001, 148,
A1225.
557 N. Yabuuchi, M. Kajiyama, J. Iwatate, H. Nishikawa, S.
Hitomi, R. Okuyama, R. Usui, Y. Yamada and S. Komaba,
Nat. Mater., 2012, 11, 512–517.
558 X. Wang, G. Liu, T. Iwao, M. Okubo and A. Yamada, J. Phys.
Chem. C, 2014, 118, 2970–2976.
559 T. Ohzuku and Y. Makimura, Chem. Lett, 2001, 30, 642–643.
560 H.-J. Noh, S. Youn, C. S. Yoon and Y.-K. Sun, J. Power
Sources, 2013, 233, 121–130.
561 C. Zhao, F. Ding, Y. Lu, L. Chen and Y.-S. Hu, Angew.
Chem., Int. Ed., 2020, 59, 264–269.
562 T. Chen, W. Liu, Y. Zhuo, H. Hu, M. Zhu, R. Cai, X. Chen,
J. Yan and K. Liu, J. Energy Chem., 2020, 43, 148–154.
563 Y. Zheng, J. Li, S. Ji, K. S. Hui, S. Wang, H. Xu, K. Wang,
D. A. Dinh, C. Zha, Z. Shao and K. N. Hui, Small, 2023,
15, 2302160.
564 T. Jin, H. Li, K. Zhu, P.-F. Wang, P. Liu and L. Jiao, Chem.
Soc. Rev., 2020, 49, 2342–2377.
565 B. Senthilkumar, C. Murugesan, L. Sharma, S. Lochab and
P. Barpanda, Small Methods, 2019, 3, 1800253.
566 S. K. Sapra, J. Pati, P. K. Dwivedi, S. Basu, J.-K. Chang and
R. S. Dhaka, Wiley Interdiscip. Rev.: Energy Environ., 2021,
10, e400.
567 S.-J. Lim, D.-W. Han, D.-H. Nam, K.-S. Hong, J.-Y. Eom,
W.-H. Ryu and H.-S. Kwon, J. Mater. Chem. A, 2014, 2,
19623–19632.
568 W. Shen, H. Li, Z. Guo, Z. Li, Q. Xu, H. Liu and Y. Wang,
RSC Adv., 2016, 6, 71581–71588.
569 L. Li, Y. Qin, S. Zhang, H. Zhao, J. Zhao, X. Li, J. Zhao,
H. Wu, Y. Su and S. Ding, J. Power Sources, 2023,
576, 233226.
570 J.-Z. Guo, H.-X. Zhang, Z.-Y. Gu, M. Du, H.-Y. Lu
¨,
X.-X. Zhao, J.-L. Yang, W.-H. Li, S. Kang, W. Zou and
X.-L. Wu, Adv. Funct. Mater., 2022, 32, 2209482.
571 M. Chen, L. Chen, Z. Hu, Q. Liu, B. Zhang, Y. Hu, Q. Gu,
J.-L. Wang, L.-Z. Wang, X. Guo, S.-L. Chou and S.-X. Dou,
Adv. Mater., 2017, 29, 1605535.
572 A. Chiring and P. Senguttuvan, Bull. Mater. Sci., 2020,
43, 306.
573 C.-T. Chu, A. Mondal, N. V. Kosova and J.-Y. Lin, Appl. Surf.
Sci., 2020, 530, 147169.
574 M. M. Thackeray and K. Amine, Nat. Energy., 2021, 6, 566.
575 Q. Zhong, A. Bonakdarpour, M. Zhang, Y. Gao and
J. R. Dahn, J. Electrochem. Soc., 1997, 144, 205.
576 W. Zhu, A. Li, Z. Wang, J. Yang and Y. Xu, Small, 2021,
17, 2006424.
577 A. Indra, T. Song and U. Paik, Adv. Mater., 2018,
30, 1705146.
578 Y. He, B. Li, M. O’Keeffe and B. Chen, Chem. Soc. Rev.,
2014, 43, 5618–5656.
579 Z. Peng, X. Yi, Z. Liu, J. Shang and D. Wang, ACS Appl.
Mater. Interfaces, 2016, 8, 14578–14585.
580 B. Huang, Y. Liu, Z. Lu, M. Shen, J. Zhou, J. Ren, X. Li and
S. Liao, ACS Sustain. Chem. Eng., 2019, 7, 16659–16667.
581 Y. Ma, Y. Ma, S. L. Dreyer, Q. Wang, K. Wang, D.
Goonetilleke, A. Omar, D. Mikhailova, H. Hahn, B.
Breitung and T. Brezesinski, Adv. Mater., 2021, 33,
2101342.
582 K. Wada, K. Sakaushi, S. Sasaki and H. Nishihara, Angew.
Chem., Int. Ed., 2018, 57, 8886–8890.
Chem Soc Rev Review Article
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2024
583 Q. Jiang, P. Xiong, J. Liu, Z. Xie, Q. Wang, X.-Q. Yang,
E. Hu, Y. Cao, J. Sun, Y. Xu and L. Chen, Angew. Chem., Int.
Ed., 2020, 59, 5273–5277.
584 A. P. Co
ˆte
´, A. I. Benin, N. W. Ockwig, M. O’Keeffe,
A. J. Matzger and O. M. Yaghi, Science, 2005, 310,
1166–1170.
585 S. Wang, Q. Wang, P. Shao, Y. Han, X. Gao, L. Ma, S. Yuan,
X. Ma, J. Zhou, X. Feng and B. Wang, J. Am. Chem. Soc.,
2017, 139, 4258–4261.
586 H. Li, M. Tang, Y. Wu, Y. Chen, S. Zhu, B. Wang, C. Jiang,
E. Wang and C. Wang, J. Phys. Chem. Lett., 2018, 9,
3205–3211.
587 C. Zhu, G. Fang, J. Zhou, J. Guo, Z. Wang, C. Wang, J. Li,
Y. Tang and S. Liang, J. Mater. Chem. A, 2018, 6, 9677–9683.
588 J.-S. M. Lee and A. I. Cooper, Chem. Rev., 2020, 120,
2171–2214.
589 Y. Liu, G. He, H. Jiang, I. P. Parkin, P. R. Shearing and
D. J. L. Brett, Adv. Funct. Mater., 2021, 31, 2010445.
590 K. W. Nam, S. Kim, S. Lee, M. Salama, I. Shterenberg,
Y. Gofer, J.-S. Kim, E. Yang, C. S. Park, J.-S. Kim, S.-S. Lee,
W.-S. Chang, S.-G. Doo, Y. N. Jo, Y. Jung, D. Aurbach and
J. W. Choi, Nano Lett., 2015, 15, 4071–4079.
591 H. Li, H. Yang, Z. Sun, Y. Shi, H.-M. Cheng and F. Li, Nano
Energy, 2019, 56, 100–108.
592 M. A. Pope and I. A. Aksay, Adv. Energy Mater, 2015,
5, 1500124.
593 W.-P. Wang, J. Zhang, J. Chou, Y.-X. Yin, Y. You, S. Xin and
Y.-G. Guo, Adv. Energy Mater., 2021, 11, 2000791.
594 C. Fu, M. B. Oviedo, Y. Zhu, A. von Wald Cresce,
K. Xu, G. Li, M. E. Itkis, R. C. Haddon, M. Chi, Y. Han,
B. M. Wong and J. Guo, ACS Nano, 2018, 12, 9775–9784.
595 W. C. Du, J. Zhang, Y. X. Yin, Y. G. Guo and L. J. Wan, Asian
J. Chem., 2016, 11, 2690–2694.
596 C. Wu, Y. Lei, L. Simonelli, D. Tonti, A. Black, X. Lu,
W. H. Lai, X. Cai, Y. X. Wang, Q. Gu, S. L. Chou,
H. K. Liu, G. Wang and S. X. Dou, Adv. Mater., 2022,
34, e2108363.
597 D. Su, M. Cortie, H. Fan and G. Wang, Adv. Mater., 2017,
29, 1700587.
598 S. Xin, Y.-X. Yin, Y.-G. Guo and L.-J. Wan, Adv. Mater., 2014,
26,1261–1265.
599 P. V. C. Medeiros, S. Marks, J. M. Wynn, A. Vasylenko,
Q. M. Ramasse, D. Quigley, J. Sloan and A. J. Morris, ACS
Nano, 2017, 11, 6178–6185.
600 W.-C. Du, J. Zhang, Y.-X. Yin, Y.-G. Guo and L.-J. Wan,
Chem. Asian J., 2016, 11, 2690–2694.
601 A. Kwade, W. Haselrieder, R. Leithoff, A. Modlinger,
F. Dietrich and K. Droeder, Nat. Energy, 2018, 3,
290–300.
602 F. Xiao, H. Wang, J. Xu, W. Yang, X. Yang, D. Y. W. Yu and
A. L. Rogach, Adv. Energy Mater., 2021, 11, 2100989.
603 C. Ye, Y. Jiao, D. Chao, T. Ling, J. Shan, B. Zhang, Q. Gu,
K. Davey, H. Wang and S.-Z. Qiao, Adv. Mater., 2020,
32, 1907557.
604 C. Wu, Y. Lei, L. Simonelli, D. Tonti, A. Black, C. Marini,
X. Lu, W.-H. Lai, X. Cai, Y.-X. Wang, Q. Gu, S.-L. Chou,
H.-K. Liu, G. Wang and S.-X. Dou, Adv. Mater., 2022,
34, 2205634.
605 J. Kim, J. H. Kim and K. Ariga, Joule, 2017, 1, 739–768.
606 H. Kim, J. Lee, H. Ahn, O. Kim and M. J. Park, Nat.
Commun., 2015, 6, 7278.
607 L. F. Zhao, Z. Hu, W. H. Lai, Y. Tao, J. Peng, Z. C. Miao,
Y. X. Wang, S. L. Chou, H. K. Liu and S. X. Dou, Adv. Energy
Mater., 2021, 11, 2002704.
608 Y. Li, Y. Lu, Q. Meng, A. C. S. Jensen, Q. Zhang, Q. Zhang,
Y. Tong, Y. Qi, L. Gu, M.-M. Titirici and Y.-S. Hu, Adv.
Energy Mater., 2019, 9, 1902852.
609 D. Li, X. Ren, Q. Ai, Q. Sun, L. Zhu, Y. Liu, Z. Liang,
R. Peng, P. Si, J. Lou, J. Feng and L. Ci, Adv. Energy Mater.,
2018, 8, 1802386.
610 X. Yao, Y. Ke, W. Ren, X. Wang, F. Xiong, W. Yang, M. Qin,
Q. Li and L. Mai, Adv. Energy Mater, 2019, 9, 1803260.
611 J. An, H. Y. Zhang, L. Qi, G. X. Li and Y. L. Li, Angew. Chem.,
Int. Ed., 2022, 134, e202113313.
612 W. Liu, H. Zhang, W. Ye, B. Xiao, Z. Sun, Y. Cheng and
M.-S. Wang, Small, 2023, 19, 2300605.
613 Y. Zheng, Y. Lu, X. Qi, Y. Wang, L. Mu, Y. Li, Q. Ma,
J. Li and Y.-S. Hu, Energy Storage Mater., 2019, 18,
269–279.
614 H. Wang, J.-L. Lan, H. Yuan, S. Luo, Y. Huang, Y. Yu, Q. Cai
and X. Yang, Appl. Surf. Sci., 2020, 518,146221.
615 C. Yang, X. Sun, X. Zhang, J. Li, J. Ma, Y. Li, L. Xu, S. Liu,
J. Yang, S. Fang, Q. Li, X. Yang, F. Pan, J. Lu and D. Yu,
Carbon, 2021, 176, 242–252.
616 C. Yang, X. Zhang, J. Li, J. Ma, L. Xu, J. Yang, S. Liu,
S. Fang, Y. Li, X. Sun, X. Yang, F. Pan, J. Lu and D. Yu,
Electrochim. Acta, 2020, 346, 136244.
617 Y. C. Lee and S. C. Jung, Nanoscale Adv., 2022, 4,
5378–5391.
618 L. Zhao, T. Tang, W. Chen, X. Feng and L. Mi, Green Energy
Environ, 2018, 3, 277–285.
619 J. Yang, X. Xiao, W. Gong, L. Zhao, G. Li, K. Jiang, R. Ma,
M. H. Rummeli, F. Li, T. Sasaki and F. Geng, Angew. Chem.,
Int. Ed., 2019, 58, 8740–8745.
620 L. Zhang, Y. Pan, Y. Chen, M. Li, P. Liu, C. Wang, P. Wang
and H. Lu, Chem. comm., 2019, 55, 4258–4261.
621 B. Wang, E. H. Ang, Y. Yang, Y. Zhang, H. Geng, M. Ye and
C. C. Li, Adv. Funct. Mater., 2020, 30, 2001708.
622 Y. Wang, M. Liu, Z. Wang, Q. Gu, B. Liu, C. Zhao, J. Zhang,
S. Xu, M. Lu, H. Li and B. Zhang, J. Energy Chem., 2022, 68,
306–313.
623 C. Xue, Y. Zhang, Z. Nie, C. Du, J. Zhang and J. Zhang,
Electrochim. Acta, 2023, 444, 142021.
624 L. Xu, X. Chen, W. Guo, L. Zeng, T. Yang, P. Xiong,
Q. Chen, J. Zhang, M. Wei and Q. Qian, Nanoscale, 2021,
13, 5033–5044.
625 J. Liao, Q. Hu, J. Mu, F. Chen, X. He, F. Chen, Z. Wen and
C. Chen, Chem. comm., 2020, 56, 8392–8395.
626 H. Zhang, Y. Geng, J. Huang, Z. Wang, K. Du and H. Li,
Energy Environ. Sci., 2023, 16, 889–951.
627 W. Zhang, H. Zuo, Z. Cheng, Y. Shi, Z. Guo, N. Meng,
A. Thomas and Y. Liao, Adv. Mater., 2022, 34, 2104952.
Review Article Chem Soc Rev
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
This journal is © The Royal Society of Chemistry 2024 Chem. Soc. Rev.
628 L. Zhong, J. Li, C. Liu, L. Fang, Z. Yuan, D. Yu and X. Chen,
Adv. Funct. Mater., 2023, 33, 2215133.
629 L. Meng, S. Ren, C. Ma, Y. Yu, Y. Lou, D. Zhang and Z. Shi,
Chem. comm., 2019, 55, 9491–9494.
630 X. Zhang, F. Wu, X. Lv, L. Xu, R. Huang, R. Chen and L. Li,
Adv. Mater., 2022, 34, 2204370.
631 S. Qiao, Q. Zhou, M. Ma, H. K. Liu, S. X. Dou and S. Chong,
ACS Nano, 2023, 17(12), 11220–11252.
632 C. Yan, C. Lv, Y. Zhu, G. Chen, J. Sun and G. Yu, Adv.
Mater., 2017, 29, 1703909.
633 C. Park, Y. H. Kim, H. Lee, H. S. Kang, T. Kim, S. W. Lee,
K. Lee, K.-B. Kim and C. Park, Adv. Energy Mater., 2021,
11, 2003243.
634 H. Li, S. Hao, Z. Tian, Z. Zhao and X. Wang, Electrochim.
Acta, 2019, 321, 134624.
635 W. Zhang, W. Li and X. Zhou, Batter. Supercaps, 2020, 3,
439–449.
636 Y. Xi, D. Yang, H. Lou, Y. Gong, C. Yi, G. Lyu, W. Han,
F. Kong and X. Qiu, Ind. Crops Prod., 2021, 161, 113179.
637 M. Peng, K. Shin, L. Jiang, Y. Jin, K. Zeng, X. Zhou and
Y. Tang, Angew. Chem., Int. Ed., 2022, 61, e202206770.
638 Y. Qiuran, F. Qining, P. Jian, C. Shulei, L. Huakun and
W. Jiazhao, Microstructures, 2023, 3, 2023013.
639 W. Xueting, W. Yunchuang, W. Meichao, F. Ruopian, Y. Xi
and W. Da-Wei, Microstructures, 2022, 2, 2022016.
640 Q. Ai, Q. Fang, J. Liang, X. Xu, T. Zhai, G. Gao, H. Guo,
G. Han, L. Ci and J. Lou, Nano Energy, 2020, 72, 104657.
641 Z. Yan, L. Zhao, Y. Wang, Z. Zhu and S. L. Chou, Adv. Funct.
Mater., 2022, 32, 2205622.
642 H. Sun, G. Zhu, X. Xu, M. Liao, Y. Y. Li, M. Angell, M. Gu,
Y. Zhu, W. H. Hung, J. Li, Y. Kuang, Y. Meng, M. C. Lin,
H. Peng and H. Dai, Nat. Commun., 2019, 10, 3302.
643 H. Duan, Y. You, G. Wang, X. Ou, J. Wen, Q. Huang, P. Lyu,
Y. Liang, Q. Li, J. Huang, Y.-X. Wang, H.-K. Liu, S. X. Dou
and W.-H. Lai, Nano-Micro Lett., 2024, 16, 78.
644 L. Zhao, Y. Tao, W. H. Lai, Z. Hu, J. Peng, Y. Lei, Y. Cao,
S. L. Chou, Y. X. Wang, H. K. Liu and S. X. Dou, Adv. Funct.
Mater., 2024, 34, 2302026.
645 L. F. Zhao, Z. Hu, Z. Y. Huang, Y. Tao, W. H. Lai, A. L. Zhao,
Q. N. Liu, J. Peng, Y. J. Lei, Y. X. Wang, Y. L. Cao, C. Wu,
S. L. Chou, H. K. Liu and S. X. Dou, Adv. Energy Mater.,
2022, 12, 2200990.
646 X. Li, Y. Tian, L. Shen, Z. Qu, T. Ma, F. Sun, X. Liu,
C. Zhang, J. Shen, X. Li, L. Gao, S. Xiao, T. Liu, Y. Liu
and Y. Lu, Adv. Funct. Mater., 2021, 31, 2009718.
647 J. Yun, B.-K. Park, E.-S. Won, S. H. Choi, H. C. Kang,
J. H. Kim, M.-S. Park and J.-W. Lee, ACS Energy Lett.,
2020, 5, 3108–3114.
648 D. Chen, S. Huang, L. Zhong, S. Wang, M. Xiao, D. Han
and Y. Meng, Adv. Funct. Mater., 2020, 30, 1907717.
649 J. He, A. Bhargav and A. Manthiram, Angew. Chem., Int. Ed.,
2022, 61, e202116586.
650 M. Chen, Q. Liu, S. W. Wang, E. Wang, X. Guo and
S. L. Chou, Adv. Energy Mater, 2019, 9, 1803609.
651 J. He, A. Bhargav and A. Manthiram, Angew. Chem., Int. Ed.,
2022, 61, e202116586.
652 S. Zhou, W. Chen, J. Shi, G. Li, F. Pei, S. Liu, W. Ye, L.
Xiao, M.-S. Wang, D. Wang, Y. Qiao, L. Huang, G.-L. Xu,
H.-G. Liao, J.-F. Chen, K. Amine and S.-G. Sun, Energy
Environ. Sci., 2022, 15, 196–205.
653 H. R. Shin, J. Yun, G. H. Eom, J. Moon, J. H. Kim, M.-S.
Park, J.-W. Lee and S. X. Dou, Nano Energy, 2022, 95, 106999.
654 S. A. Han, H. Qutaish, J.-W. Lee, M.-S. Park and J. H. Kim,
EcoMat, 2023, 5, e12283.
655 H. R. Shin, S. Kim, J. Park, J. H. Kim, M.-S. Park and
J.-W. Lee, Energy Storage Mater., 2023, 56, 515–523.
656 J. Kim, J. Lee, J. Yun, S. H. Choi, S. A. Han, J. Moon,
J. H. Kim, J.-W. Lee and M.-S. Park, Adv. Funct. Mater.,
2020, 30, 1910538.
657 R. Zhao, J. Yang, X. Han, Y. Wang, Q. Ni, Z. Hu, C. Wu and
Y. Bai, Adv. Energy Mater., 2023, 13, 2203542.
658 H. Liu, Q. Ye, D. Lei, Z. Hou, W. Hua, Y. Huyan, N. Li,
C. Wei, F. Kang and J.-G. Wang, Energy Environ. Sci., 2023,
16, 1610–1619.
659 L. Yuan, J. Hao, C.-C. Kao, C. Wu, H.-K. Liu, S.-X. Dou and
S.-Z. Qiao, Energy Environ. Sci., 2021, 14, 5669–5689.
660 K. Amin, N. Ashraf, L. Mao, C. F. J. Faul and Z. Wei, Nano
Energy, 2021, 85, 105958.
661 Y. Sun, J. Tang, K. Zhang, X. Yu, J. Yuan, D.-M. Zhu,
K. Ozawa and L.-C. Qin, RSC Adv.,2021, 11, 34152–34159.
662 Y. Yang, C. Zhang, Z. Mei, Y. Sun, Q. An, Q. Jing, G. Zhao
and H. Guo, Nano Res., 2023, 16(7), 9289–9298.
663 B. Zhou, A. Hu, X. Zeng, M. He, R. Li, C. Zhao, Z. Yan,
Y. Pan, J. Chen, Y. Fan, M. Liu and J. Long, J. Chem. Eng.,
2022, 450, 137921.
664 S. You, D.-S. Liu, M. Ye, Y. Zhang, Y. Tang, X. Liu and
C. Chao Li, J. Chem. Eng, 2023, 454, 139907.
665 P. Jaumaux, J. Wu, D. Shanmukaraj, Y. Wang, D. Zhou,
B. Sun, F. Kang, B. Li, M. Armand and G. Wang, Adv. Funct.
Mater., 2021, 31, 2008644.
666 Q. Zhao, S. Stalin, C.-Z. Zhao and L. A. Archer, Nat. Rev.
Mater., 2020, 5, 229–252.
667 J. Lu, P. Jaumaux, T. Wang, C. Wang and G. Wang, J. Mater.
Chem. A, 2021, 9, 24175–24194.
668 Y. S. Meng, V. Srinivasan and K. Xu, Science, 2022,
378, eabq3750.
669 J. Li, J. K. Park, R. B. Moore and L. A. Madsen, Nat. Mater.,
2011, 10, 507–511.
670 Q. Wen, D. Yan, F. Liu, M. Wang, Y. Ling, P. Wang,
P. Kluth, D. Schauries, C. Trautmann and P. Apel, Adv.
Funct. Mater., 2016, 26, 5796–5803.
671 C. L. Ritt, M. Liu, T. A. Pham, R. Epsztein, H. J. Kulik and
M. Elimelech, Sci. Adv., 2022, 8, eabl5771.
672 H. Khani, S. Kalami and J. B. Goodenough, Sustainable
Energy Fuels, 2020, 4, 177–189.
673 H. Zou, Y. Wang, X. Li, P. Ding, H. Guo and F. Li, J. Membr.
Sci., 2023, 685, 121939.
674 D.-D. Han, Z.-Y. Wang, G.-L. Pan and X.-P. Gao, ACS Appl.
Mater. Interfaces, 2019, 11, 18427–18435.
675 Z. Wu, Y. Yi, F. Hai, X. Tian, S. Zheng, J. Guo, W. Tang,
W. Hua and M. Li, ACS Appl. Mater. Interfaces, 2023, 15,
22065–22074.
Chem Soc Rev Review Article
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2024
676 J.Y.Jung,S.A.Han,H.-S.Kim,J.H.Suh,J.-S.Yu,W.Cho,
M.-S. Park and J. H. Kim, ACS Nano, 2023, 17, 15931–15941.
677 Z. Li, L. Wang, Y. Liu, M. Yu, B. Liu, Y. Men, Z. Sun, W. Hu
and G. Zhu, Small, 2023, 19, 2302818.
678 Z. Guo, Y. Zhang, Y. Dong, J. Li, S. Li, P. Shao, X. Feng and
B. Wang, J. Am. Chem. Soc., 2019, 141, 1923–1927.
679 S. Lascaud, M. Perrier, A. Vallee, S. Besner, J. Prud’homme
and M. Armand, Macromol., 1994, 27, 7469–7477.
680 W. Gorecki, M. Jeannin, E. Belorizky, C. Roux and
M. Armand, J. Condens. Matter Phys., 1995, 7, 6823.
681 Z. Stoeva, I. Martin-Litas, E. Staunton, Y. G. Andreev and
P. G. Bruce, J. Am. Chem. Soc., 2003, 125, 4619–4626.
682 Y. Nikodimos, W. N. Su and B. J. Hwang, Adv. Energy
Mater., 2023, 13, 2202854.
683 W. Zhao, J. Yi, P. He and H. Zhou, Electrochem. Energy Rev.,
2019, 2, 574–605.
684 S. Yu, J. Noh, B. Kim, J.-H. Song, K. Oh, J. Yoo, S. Lee,
S.-O. Park, W. Kim, B. Kang, D. Kil and K. Kang, Science,
2023, 382, 573–579.
685 B. Zhang, R. Tan, L. Yang, J. Zheng, K. Zhang, S. Mo, Z. Lin
and F. Pan, Energy Storage Mater, 2018, 10, 139–159.
686 X. Feng, H. Fang, N. Wu, P. Liu, P. Jena, J. Nanda and
D. Mitlin, Joule, 2022, 6, 543–587.
687 X. Yu and A. Manthiram, Energy Environ. Sci., 2018, 11,
527–543.
688 Y. Huang, B. Shao and F. Han, Curr. Opin. Electrochem.,
2022, 33, 100933.
689 S. Lou, F. Zhang, C. Fu, M. Chen, Y. Ma, G. Yin and
J. Wang, Adv. Mater., 2021, 33, 2000721.
690 Y. Huang, L. Zhao, L. Li, M. Xie, F. Wu and R. Chen, Adv.
Mater., 2019, 31, 1808393.
691 N. Ohta, K. Takada, L. Zhang, R. Ma, M. Osada and
T. Sasaki, Adv. Mater., 2006, 18, 2226–2229.
692 Y. Li, W. Zhou, X. Chen, X. Lu
¨, Z. Cui, S. Xin, L. Xue, Q. Jia
and J. B. Goodenough, Proc. Natl. Acad. Sci. U. S. A., 2016,
113, 13313–13317.
693 K. Yamamoto, R. Yoshida, T. Sato, H. Matsumoto, H.
Kurobe, T. Hamanaka, T. Kato, Y. Iriyama and T.
Hirayama, J. Power Sources, 2014, 266, 414–421.
694 S. Bai, B. Kim, C. Kim, O. Tamwattana, H. Park, J. Kim, D. Lee
and K. Kang, Nat. Nanotechnol., 2021, 16,7784.
695 H. Wu, D. Zhuo, D. Kong and Y. Cui, Nat. Commun., 2014,
5, 5193.
696 M. F. Lagadec, R. Zahn and V. Wood, Nat. Energy., 2019, 4,
16–25.
697 J. Pu, J. Li, K. Zhang, T. Zhang, C. Li, H. Ma, J. Zhu,
P. V. Braun, J. Lu and H. Zhang, Nat. Commun., 2019, 10,
1896.
698 Z. Liu, Y. Jiang, Q. Hu, S. Guo, L. Yu, Q. Li, Q. Liu and
X. Hu, Energy Environ. Mater., 2021, 4, 336–362.
699 L. Zhang, X. Li, M. Yang and W. Chen, Energy Storage
Mater., 2021, 41, 522–545.
700 W. Zhang, Z. Tu, J. Qian, S. Choudhury, L. A. Archer and
Y. Lu, Small, 2018, 14, 1703001.
701 P. Sun, K. Wang and H. Zhu, Adv. Mater., 2016, 28,
2287–2310.
702 Y. Cao, C. Liu, M. Wang, H. Yang, S. Liu, H. Wang, Z. Yang,
F. Pan, Z. Jiang and J. Sun, Energy Storage Mater, 2020, 29,
207–215.
703 J. Xu, H. Zhang, F. Yu, Y. Cao, M. Liao, X. Dong and
Y. Wang, Angew. Chem., Int. Ed., 2022, 61, e202211933.
704 K. Mullapudi, R. Addou, C. L. Dezelah, D. F. Moser,
R. K. Kanjolia, J. H. Woodruff and J. F. ConleyJr., Chem.
Mater., 2023, 35, 4649–4659.
705 Z. Zhao, W. Chen, S. Impeng, M. Li, R. Wang, Y. Liu,
L. Zhang, L. Dong, J. Unruangsri, C. Peng, C. Wang,
S. Namuangruk, S.-Y. Lee, Y. Wang, H. Lu and J. Guo,
J. Mater. Chem. A, 2020, 8, 3459–3467.
706 J. Li, C. Jiao, J. Zhu, L. Zhong, T. Kang, S. Aslam, J. Wang,
S. Zhao and Y. Qiu, J. Energy Chem., 2021, 57, 469–476.
707 C. Kisielowski, B. Freitag, M. Bischoff, H. van Lin, S. Lazar,
G. Knippels, P. Tiemeijer, M. van der Stam, S. von Harrach,
M. Stekelenburg, M. Haider, S. Uhlemann, H. Mu
¨ller,
P. Hartel, B. Kabius, D. Miller, I. Petrov, E. A. Olson,
T. Donchev, E. A. Kenik, A. R. Lupini, J. Bentley, S. J.
Pennycook, I. M. Anderson, A. M. Minor, A. K. Schmid,
T. Duden, V. Radmilovic, Q. M. Ramasse, M. Watanabe,
R. Erni, E. A. Stach, P. Denes and U. Dahmen, Microsc.
Microanal., 2008, 14, 469–477.
708 S. I. Wright, M. M. Nowell and D. P. Field, Microsc.
Microanal.,2011, 17, 316–329.
709 Y. Li, Y. Li, A. Pei, K. Yan, Y. Sun, C.-L. Wu, L.-M. Joubert,
R. Chin, A. L. Koh, Y. Yu, J. Perrino, B. Butz, S. Chu and
Y. Cui, Science, 2017, 358, 506–510.
710 X. Wang, Y. Li and Y. S. Meng, Joule, 2018, 2, 2225–2234.
711 S. Dolabella, A. Borzı
`, A. Dommann and A. Neels, Small
Methods, 2022, 6, 2100932.
712 Y. Umena, K. Kawakami, J.-R. Shen and N. Kamiya, Nature,
2011, 473, 55–60.
713 Y. Fukuda, Y. Hirano, K. Kusaka, T. Inoue and T. Tamada,
Proc. Natl. Acad. Sci. U. S. A., 2020, 117, 4071–4077.
714 S. Mahapatra, L. Li, J. F. Schultz and N. Jiang, J. Chem.
Phys., 2020, 153, 010902.
715 Z. Wu, W. Kong Pang, L. Chen, B. Johannessen and Z. Guo,
Batter. Supercaps, 2021, 4, 1547–1566.
716 J. Kim, D. Lee, C. Nam, J. Chung, B. Koo, N. Kim and J. Lim,
J. Electron. Spectrosc. Relat. Phenom., 2023, 266, 147337.
717 Y. Yan, C. Cheng, L. Zhang, Y. Li and J. Lu, Adv. Energy
Mater., 2019, 9, 1900148.
718 Y. S. Meng and M. E. Arroyo-de Dompablo, Energy Environ.
Sci., 2009, 2, 589–609.
719 S. Bai, Y. Sun, J. Yi, Y. He, Y. Qiao and H. Zhou, Joule, 2018,
2, 2117–2132.
720 L. Wang, F. Zhou, Y. S. Meng and G. Ceder, Phys. Rev. B,
2007, 76, 165435.
721 T. Lombardo, M. Duquesnoy, H. El-Bouysidy, F. Åre
´n, A.
Gallo-Bueno, P. B. Jørgensen, A. Bhowmik, A. Demortie
`re,
E. Ayerbe, F. Alcaide, M. Reynaud, J. Carrasco, A. Grimaud,
C. Zhang, T. Vegge, P. Johansson and A. A. Franco, Chem.
Rev., 2022, 122, 10899–10969.
722 Y. Xiao, Y. Wang, S.-H. Bo, J. C. Kim, L. J. Miara and
G. Ceder, Nat. Rev. Mater., 2020, 5, 105–126.
Review Article Chem Soc Rev
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
This journal is © The Royal Society of Chemistry 2024 Chem. Soc. Rev.
723 H.-K. Tian, R. Jalem, B. Gao, Y. Yamamoto, S. Muto,
M. Sakakura, Y. Iriyama and Y. Tateyama, ACS Appl. Mater.
Interfaces, 2020, 12, 54752–54762.
724 Z. Sun, K. Xi, J. Chen, A. Abdelkader, M.-Y. Li, Y. Qin,
Y. Lin, Q. Jiang, Y.-Q. Su, R. Vasant Kumar and S. Ding,
Nat. Commun., 2022, 13, 3209.
725Y.Gao,T.Rojas,K.Wang,S.Liu,D.Wang,T.Chen,H.
Wang,A.T.NgoandD.Wang,Nat. Energy., 2020, 5, 534–542.
726 Z. Yan, L. Zhao, Y. Wang, Z. Zhu and S.-L. Chou, Adv. Funct.
Mater., 2022, 32, 2205622.
727 H. Ye, M. Li, T. Liu, Y. Li and J. Lu, ACS Energy Lett., 2020,
5, 2234–2245.
728 Y. Kim, N. J. Dudney, M. Chi, S. K. Martha, J. Nanda,
G. M. Veith and C. Liang, J. Electrochem. Soc., 2013,
160, A3113.
729 S. Rajendran, A. George, Z. Tang, C. Neumann, A. Turchanin
andL.M.R.Arava,Small, 2023, 19, 2303625.
730 B. Campbell, R. Ionescu, Z. Favors, C. S. Ozkan and
M. Ozkan, Sci. Rep., 2015, 5, 14575.
731 A. Karatrantos and Q. Cai, Phys. Chem. Chem. Phys., 2016,
18, 30761–30769.
732 S. Sun, Z. Han, W. Liu, Q. Xia, L. Xue, X. Lei, T. Zhai, D. Su
and H. Xia, Nat. Commun., 2023, 14, 6662.
733 Q. Li, Y. Qiao, S. Guo, K. Jiang, Q. Li, J. Wu and H. Zhou,
Joule, 2018, 2, 1134–1145.
734 Y.-J. Lei, H.-L. Yang, Y. Liang, H.-W. Liu, B. Zhang,
L. Wang, W.-H. Lai, Y.-X. Wang, H.-K. Liu and S.-X. Dou,
Adv. Energy Mater., 2022, 12, 2202523.
735 N. An-Giang, V. Rakesh, N. D. Pravin and P. Chan-Jin,
Energy Materials, 2023, 3, 300030.
736 R. Ding, Y. Huang, G. Li, Q. Liao, T. Wei, Y. Liu, Y. Huang
and H. He, Front. Chem., 2020, 8, 607504.
737 R. J. Cle
´ment, P. G. Bruce and C. P. Grey, J. Electrochem.
Soc., 2015, 162, A2589.
Chem Soc Rev Review Article
Published on 04 March 2024. Downloaded by University of Wollongong on 3/8/2024 12:31:15 AM.
View Article Online
Article
Full-text available
The concentration difference in the near-surface region of lithium metal is the main cause of lithium dendrite growth. Resolving this issue will be key to achieving high-performance lithium metal batteries (LMBs). Herein, we construct a lithium nitrate (LiNO3)-implanted electroactive β phase polyvinylidene fluoride-co-hexafluoropropylene (PVDF-HFP) crystalline polymorph layer (PHL). The electronegatively charged polymer chains attain lithium ions on the surface to form lithium-ion charged channels. These channels act as reservoirs to sustainably release Li ions to recompense the ionic flux of electrolytes, decreasing the growth of lithium dendrites. The stretched molecular channels can also accelerate the transport of Li ions. The combined effects enable a high Coulombic efficiency of 97.0% for 250 cycles in lithium (Li)||copper (Cu) cell and a stable symmetric plating/stripping behavior over 2000 h at 3 mA cm⁻² with ultrahigh Li utilization of 50%. Furthermore, the full cell coupled with PHL-Cu@Li anode and LiFePO4 cathode exhibits long-term cycle stability with high-capacity retention of 95.9% after 900 cycles. Impressively, the full cell paired with LiNi0.87Co0.1Mn0.03O2 maintains a discharge capacity of 170.0 mAh g⁻¹ with a capacity retention of 84.3% after 100 cycles even under harsh condition of ultralow N/P ratio of 0.83. This facile strategy will widen the potential application of LiNO3 in ester-based electrolyte for practical high-voltage LMBs. Supplementary Information The online version contains supplementary material available at 10.1007/s40820-023-01300-5.
Article
Full-text available
Large lattice expansion/contraction with Li⁺ intercalation/deintercalation of electrode active materials results in severe structural degradation to electrodes and can negatively impact the cycle life of solid-state lithium-based batteries. In case of the layered orthorhombic MoO3 (α-MoO3), its large lattice variation along the b axis during Li⁺ insertion/extraction induces irreversible phase transition and structural degradation, leading to undesirable cycle life. Herein, we propose a lattice pinning strategy to construct a coherent interface between α-MoO3 and η-Mo4O11 with epitaxial intergrowth structure. Owing to the minimal lattice change of η-Mo4O11 during Li⁺ insertion/extraction, η-Mo4O11 domains serve as pin centers that can effectively suppress the lattice expansion of α-MoO3, evidenced by the noticeably decreased lattice expansion from about 16% to 2% along the b direction. The designed α-MoO3/η-Mo4O11 intergrown heterostructure enables robust structural stability during cycling (about 81% capacity retention after 3000 cycles at a specific current of 2 A g⁻¹ and 298 ± 2 K) by harnessing the merits of epitaxial stabilization and the pinning effect. Finally, benefiting from the stable positive electrode–solid electrolyte interface, a highly durable and flexible all-solid-state thin-film lithium microbattery is further demonstrated. This work advances the fundamental understanding of the unstable structure evolution for α-MoO3, and may offer a rational strategy to develop highly stable electrode materials for advanced batteries.
Article
Full-text available
Developing K‐ion conducting solid‐state electrolytes (SSEs) plays a critical role in the safe implementation of potassium batteries. In this work, we report a chalcogenide‐based potassium ion SSEs, K 3 SbSe 4 , which adopts a trigonal structure at room temperature. Single‐crystal structural analysis reveals a trigonal‐to‐cubic phase transition at the low temperature of 50°C, which is the lowest among similar compounds and thus provides easy access to the cubic phase. The substitution of barium for potassium in K 3 SbSe 4 leads to the creation of potassium vacancies, expansion of lattice parameters, and a transformation from a trigonal phase to a cubic phase. As a result, the maximum conductivity of K 3‐2x Ba x SbSe 4 reaches around 0.1 mS/cm at 40°C for K 2.2 Ba 0.4 SbSe 4 , which is over two orders of magnitude higher than that of undoped K 3 SbSe 4 . This novel SSE is successfully employed in a K‐O 2 battery operating at room temperature where a polymer‐laminated K 2.2 Ba 0.4 SbSe 4 pellet serves as a separator between the oxygen cathode and the potassium metal anode. Effective protection of the K metal anode against corrosion caused by O 2 has been demonstrated. This article is protected by copyright. All rights reserved
Article
Full-text available
The organic‐inorganic interfaces can enhance Li+ transport in composite solid‐state electrolytes (CSEs) due to the strong interface interactions. However, Li+ non‐conductive areas in CSEs with inert fillers will hinder the construction of efficient Li+ transport channels. Herein, CSEs with fully active Li+ conductive networks are proposed to improve Li+ transport by composing sub‐1 nm inorganic cluster chains and organic polymer chains. The inorganic cluster chains are monodispersed in polymer matrix by a brief mixed‐solvent strategy, their sub‐1 nm diameter and ultrafine dispersion state eliminate Li+ non‐conductive areas in the interior of inert fillers and filler‐agglomeration, respectively, providing rich surface areas for interface interactions. Therefore, the three‐dimensional networks connected by the monodispersed cluster chains finally construct homogeneous, large‐scale, continuous Li+ fast transport channels. Furthermore, a conjecture about 1D oriented distribution of organic polymer chains along the inorganic cluster chains is proposed to optimize Li+ pathways. Consequently, the as‐obtained CSEs possess high ionic conductivity at room temperature (0.52 mS cm−1), high Li+ transference number (0.62), and more mobile Li+ (50.7%). The assembled LiFePO4/Li cell delivers excellent stability of 1000 cycles at 0.5 C and 700 cycles at 1 C. This research provides a new strategy for enhancing Li+ transport by efficient interfaces. This article is protected by copyright. All rights reserved
Article
Full-text available
Membranes with high ion permeability and selectivity are of considerable interest for sustainable water treatment, resource extraction and energy storage. Herein, inspired by K⁺ channel of streptomyces A (KcsA K⁺), we have constructed cation sieving membranes using MXene nanosheets and Ethylenediaminetetraacetic acid (EDTA) molecules as building blocks. Numerous negatively charged oxygen atoms of EDTA molecules and 6.0 Å two-dimensional (2D) sub-nanochannel of MXene nanosheets enable biomimetic channel size, chemical groups and tunable charge density for the resulting membranes. The membranes show the capability to recognize monovalent/divalent cations, achieving excellent K⁺/Mg²⁺ selectivity of 121.2 using mixed salt solution as the feed, which outperforms other reported membranes under similar testing conditions and transcends the current upper limit. Characterization and simulations indicate that the cation recognition effect of EDTA and partial dehydration effects play critical roles in cations selective sieving and increasing the local charge density within the sub-nanochannel significantly improves cation selectivity. Our findings provide a theoretical basis for ions transport in sub-nanochannels and an alternative strategy for design ions separation membranes.
Article
To mitigate the safety issues of liquid electrolyte-based Li-ion batteries, there is a growing interest in the development of solid-state electrolyte (SSE) based Li-metal batteries. Regrettably, most SSEs have low ionic conductivity (s), which significantly lowers the performance of the Li-metal batteries. However, a composite polymer electrolyte (CPE) offers a higher s value, which still needs to be improved for a reliable Li-metal battery. Here, a superior CPE (MZ-CPE) was synthesized using modified SSZ-13 (M-SSZ-13) as a ceramic filler. Polyethylene oxide (PEO) and lithium bis-(trifluoromethanesulfonyl)imide (LiTFSI) were used as the substrate and Li-salt, respectively. Due to the upgraded and hydrophobic surface of M-SSZ-13, the dispersion of the Li-salt in PEO is significantly improved. Thus, the value of s was greatly enhanced, which helps to make better interfacial contact with the electrodes. The MZ-CPE electrolyte with 5 wt% M-SSZ-13 (5% MZ-CPE) provided an outstanding s value of 5.34 Â 10 À2 S cm À1 (@ 70 C) along with a Li-ion transference number of 0.85. Besides, the obtained discharge specific capacities were 154 and 194 mA h g À1 using LiFePO 4 and LiNiCoAlO 2 cathodes, respectively at the discharge current density of 0.1C. For LiFePO 4 cathode, the capacity retention was 94.1% after 80 cycles @ 60 C. These results indicate that hydrophobic zeolite containing composite polymer electrolytes could be a potential alternative for solid-state Li-metal batteries.
Article
Lithium-metal-halides have emerged as a class of solid electrolytes that can deliver superionic conductivity comparable to that of state-of-the-art sulfide electrolytes, as well as electrochemical stability that is suitable for high-voltage (>4 volt) operations. We show that the superionic conduction in a trigonal halide, such as Li 3 MCl 6 [where metal (M) is Y or Er], is governed by the in-plane lithium percolation paths and stacking interlayer distance. These two factors are inversely correlated with each other by the partial occupancy of M, serving as both a diffusion inhibitor and pillar for maintaining interlayer distance. These findings suggest that a critical range or ordering of M exists in trigonal halides, and we showcase the achievement of high ionic conductivity by adjusting the simple M ratio (per Cl or Li). We provide general design criteria for superionic trigonal halide electrolytes.
Article
All-solid-state sodium-ion batteries (SIBs) have the potential to offer large-scale, safe, cost-effective, and sustainable energy storage solutions by supplementing the industry-leading lithium-ion batteries. However, for the enhanced bulk properties of SIB components (e.g., solid electrolytes), a comprehensive understanding of their atomic-scale structure and the dynamic behavior of sodium (Na) ions is essential. Here, we utilize a robust multinuclear (23Na, 125Te, 25Mg, and 67Zn) magnetic resonance approach to explore a novel Mg/Zn homogeneously mixed-cation honeycomb-layered oxide Na2MgxZn2-xTeO6 solid solution series. These new intermediate compounds exhibit tailorable bulk Na-ion conductivity (σ) with the highest σ = 0.14 × 10-4 S cm-1 for Na2MgZnTeO6 at room temperature suitable for SIB solid electrolyte applications as observed by powder electrochemical impedance spectroscopy (EIS). A combination of powder X-ray diffraction (XRD), energy-dispersive X-ray (EDX) spectroscopy, and field emission scanning electron microscopy (FESEM) reveals highly crystalline phase-pure compounds in the P6322 space group. We show that the Mg/Zn disorder is random within the honeycomb layers using 125Te nuclear magnetic resonance (NMR) and resolve multiple Na sites using two-dimensional (triple-quantum magic-angle spinning (3QMAS)) 23Na NMR. The medium-range disorder in the honeycomb layer is revealed through the combination of 25Mg and 67Zn NMR, complemented by electronic structure calculations using density functional theory (DFT). Furthermore, we expose very fast local Na-ion hopping processes (hopping rate, 1/τNMR = 0.83 × 109 Hz) by using a laser to achieve variable high-temperature (∼860 K) 23Na NMR, which are sensitive to different Mg/Zn ratios. The Na2MgZnTeO6 with maximum Mg/Zn disorder displays the highest short-range Na-ion dynamics among all of the solid solution members.
Article
Organic solid electrolytes compatible with all-solid-state Li metal batteries (LMBs) are essential to ensuring battery safety, high energy density, and long-term cycling performance. However, it remains a challenge to develop an approach to provide organic solid electrolytes with capabilities for the facile dissociation of strong Li-ion pairs and fast transport of ionic components. Herein, a diethylene glycol-modified pyridinium covalent organic framework (DEG-PMCOF) with a well-defined periodic structure is prepared as a multicomponent solid electrolyte with a cationic moiety of high polarity, an additional flexible ion-transporter, and an ordered ionic channel for all-solid-state LMBs. The DEG-containing pyridinium groups of DEG-PMCOF allow a lower dissociation energy of Li salts and a smaller energy barrier of Li-ion transport, leading to high ion conductivity (1.71 × 10-4 S cm-1) and a large Li-ion transfer number (0.61) at room temperature in the solid electrolyte. The DEG-PMCOF solid electrolyte exhibits a wide electrochemical stability window and effectively suppresses the formation of Li dendrites and dead Li in all-solid-state LMBs. Molecular dynamics and density functional theory simulations provide insights into the mechanisms for the enhanced Li-ion transport driven by the integrated diffusion process based on hopping motion, vehicle motion, and free diffusion of DEG-PMCOF. The all-solid-state LMB assembled with a DEG-PMCOF solid electrolyte displays a high specific capacity with a retention of 99% and an outstanding Coulombic efficiency of 99% at various C-rates during long-term cycling. This DEG-PMCOF approach can offer an effective route to design various solid-state Li batteries.
Article
Beyond lithium-ion technologies, lithium-sulfur batteries stand out because of their multielectron redox reactions and high theoretical specific energy (2500 Wh kg-1). However, the intrinsic irreversible transformation of soluble lithium polysulfides to solid short-chain sulfur species (Li2S2 and Li2S) and the associated large volume change of electrode materials significantly impair the long-term stability of the battery. Here we present a liquid sulfur electrode consisting of lithium thiophosphate complexes dissolved in organic solvents that enable the bonding and storage of discharge reaction products without precipitation. Insights garnered from coupled spectroscopic and density functional theory studies guide the complex molecular design, complexation mechanism, and associated electrochemical reaction mechanism. With the novel complexes as cathode materials, high specific capacity (1425 mAh g-1 at 0.2 C) and excellent cycling stability (80% retention after 400 cycles at 0.5 C) are achieved at room temperature. Moreover, the highly reversible all-liquid electrochemical conversion enables excellent low-temperature battery operability (>400 mAh g-1 at -40 °C and >200 mAh g-1 at -60 °C). This work opens new avenues to design and tailor the sulfur electrode for enhanced electrochemical performance across a wide operating temperature range.