ArticlePDF Available

Electrochemo‐Mechanical Stresses and Their Measurements in Sulfide‐Based All‐Solid‐State Batteries: A Review

Wiley
Advanced Energy Materials
Authors:

Abstract and Figures

Sulfide‐based all‐solid‐state batteries (ASSBs) are one of the most promising energy storage devices due to their high energy density and good safety. However, due to the volume (stress) changes of the solid active materials during the charging and discharging process, the generation and evolution of electrochemomechanical stresses are becoming serious and unavoidable problems during the operation of all‐solid‐state batteries due to the lack of a liquid electrolyte to partially buffer the stress generated in the electrodes. To understand these electrochemo‐mechanical effects, including the origins and evolution of mechanical or internal stresses, it is necessary to develop some highly sensitive probing techniques to measure them precisely and bridge the relationship between the electrochemical reaction process and internal stress evolution. Herein, recent progress on uncovering the origins of the internal stresses, the working principle and experimental devices for stress measurement, and the application of those stress‐measuring techniques in the study of electrochemical reactions in sulfide‐based ASSBs are briefly summarized and overviewed. The investigation of precise and operando monitoring techniques and strategies for suppressing or relaxing these electrochemomechanical stresses will be an important direction in future solid‐state batteries.
a) Stack pressure evolution after pressing to an initial stack pressure of 30 MPa and holding at open circuit for two different symmetric cells based on LSPS (red) and LPSC (blue). Reproduced with permission.[14m] Copyright 2021, American Chemical Society. b) Voltage curves (red) and stack pressure profiles (blue) of two LSPS symmetric cells (operated at 0.5 mA cm⁻² (solid lines), and held at OCV (dotted lines). Reproduced with permission.[14m] Copyright 2021, American Chemical Society. c,d) Charge‐discharge voltage profiles NCM/Li ASSB and corresponding pressure changes for c) uncoated and d) In/LixIn‐coated thin LMAs. Reproduced with permission.[⁹³] Copyright 2022, Elsevier. e) Capacity‐normalized pressure change difference (Δ(ΔP)Q) as a function of cycle number. Reproduced with permission.[⁹³] Copyright 2022, Elsevier. f) Time‐resolved voltage (top), external cycling pressure (middle), and an internal optical signal (bottom) for an LTO‐LiIn ASSB cycled at C/30 (5.83 mA g⁻¹) and 25 °C in an operando mode. Reproduced with permission.[¹⁶] Copyright 2022, The Authors, published by Springer Nature. g) Stress evolution obtained from the FBG sensor placed inside Li–In anode by demodulating the optical signal (blue curve). Reproduced according to the terms of the CC‐BY license.[¹⁶] Copyright 2022, The Authors, published by Springer Nature. h) 2D stack view of the operando collected spectra by the FBG sensor placed between LTO/Li–In interface, with the corresponding galvanostatic charge/discharge cycle. A typical birefringence phenomenon could be found. Reproduced according to the terms of the CC‐BY license.[¹⁶] Copyright 2022, The Authors, published by Springer Nature. i) Galvanostatic cycle (top), λx and λy evolution (middle), and operando stress evolution obtained internally by the FBG sensor and with the experimental calibration of the sensor (bottom). Reproduced according to the terms of the CC‐BY license.[¹⁶] Copyright 2022, The Authors, published by Springer Nature. j–l) Galvanostatic cycling of j) Sb, k) Sn, and l) Si composite anodes in ASSBs with LPSC solid electrolyte and NMC‐111 cathodes, along with the measured uniaxial stress (stack pressure) changes in each cell. Reproduced with permission.[14b] Copyright 2021, Elsevier.
… 
This content is subject to copyright. Terms and conditions apply.
www.advenergymat.de
2203153 (1 of 21) © 2022 Wiley-VCH GmbH
Review
Electrochemo-Mechanical Stresses and Their
Measurements in Sulfide-Based All-Solid-State Batteries:
A Review
Jiabao Gu, Ziteng Liang, Jingwen Shi, and Yong Yang*
DOI: 10.1002/aenm.202203153
to further push the energy density of
LIBs to the limit, lithium metal is urged
as a “holy grail” anode material owing
to its high theoretical specific capacity
(3860 mAh g1) and the lowest redox
potential (3.040 V vs standard hydrogen
electrode (SHE)).[3] However, the applica-
tion of lithium metal anodes in traditional
LIBs will inevitably lead to lithium den-
drite issues, posing serious safety risks.[4]
ASSBs oer intrinsic safety advantages by
replacing liquid electrolytes with nonflam-
mable, high mechanical moduli solid-state
electrolytes (SSEs). [5] In addition, bipolar
stacking can be achieved in ASSBs, thus
minimizing inactive components in the
battery module pack and achieving higher
energy density.[6] Among various SSE
candidates (such as oxides, sulfides, poly-
mers, etc.), the sulfide solid-state electro-
lyte, with its high room temperature ionic
conductivity (103–102 S cm1) and good
formability, oers the potential for prac-
tical applications of all-solid batteries. [7]
Despite their promise, electrochemomechanical problems
in sulfide-based ASSBs remain poorly understood. These elec-
trochemomechanical eects (mainly resulting from volume
changes of active materials during cycling) in LIBs with liquid
electrolytes have already been studied. [8] However, they will
become more complex and severe when it comes to solid-state
batteries.[8a,9] Unlike the liquid battery, the internal (mechan-
ical) stress generated in the particles or electrodes in ASSBs
cannot be transferred quickly and uniformly with rigid SSEs,
thus resulting in the concentrated accumulation and release
of the internal stress of the battery.[10] Therefore, the internal
stress generated in an all-solid-state battery will lead to severe
electrochemomechanical issues such as physical contact losses
between active materials and solid electrolytes, breakdown
of the active materials, short circuits of Li metal anode, etc.[11]
These electrochemo-mechanical problems will block ion/elec-
tron transport and increase battery impedance, drastically dete-
riorating the electrochemical performance and hindering the
application of ASSBs.[10b,f,12]
Measuring the corresponding stress evolution precisely
during battery cycling is crucial for understanding the electro-
chemo-mechanical problems. To the best of our knowledge,
the earliest internal stress measurement work of sulfide-based
ASSBs was in a Li–S battery system, reported by Janek’s group
Sulfide-based all-solid-state batteries (ASSBs) are one of the most promising
energy storage devices due to their high energy density and good safety.
However, due to the volume (stress) changes of the solid active materials
during the charging and discharging process, the generation and evolution
of electrochemomechanical stresses are becoming serious and unavoidable
problems during the operation of all-solid-state batteries due to the lack of a
liquid electrolyte to partially buer the stress generated in the electrodes. To
understand these electrochemo-mechanical eects, including the origins and
evolution of mechanical or internal stresses, it is necessary to develop some
highly sensitive probing techniques to measure them precisely and bridge the
relationship between the electrochemical reaction process and internal stress
evolution. Herein, recent progress on uncovering the origins of the internal
stresses, the working principle and experimental devices for stress measure-
ment, and the application of those stress-measuring techniques in the study
of electrochemical reactions in sulfide-based ASSBs are briefly summarized
and overviewed. The investigation of precise and operando monitoring
techniques and strategies for suppressing or relaxing these electrochemome-
chanical stresses will be an important direction in future solid-state batteries.
J. Gu, Z. Liang, J. Shi, Y. Yang
State Key Laboratory for Physical Chemistry of Solid Surfaces
Collaborative Innovation Center of Chemistry for Energy Materials
and Department of Chemistry
College of Chemistry and Chemical Engineering
Xiamen University
Xiamen 361005, P. R. China
E-mail: yyang@xmu.edu.cn
Y. Yang
School of Energy Research
Xiamen University
Xiamen 361005, P. R. China
The ORCID identification number(s) for the author(s) of this article
can be found under https://doi.org/10.1002/aenm.202203153.
1. Introduction
Lithium-ion batteries (LIBs) have been widely used in consumer
electronics, electric vehicles (EVs), and grid energy storage sys-
tems since commercialized by Sony in 1991.[1] Despite great
progress in lithium-ion battery research over the past decades,
thermal runaway risks under abusive conditions due to flam-
mable organic electrolytes remain challenging.[2] Furthermore,
Adv. Energy Mater. 2022, 2203153
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202203153 by Xiamen University, Wiley Online Library on [27/11/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advenergymat.de
© 2022 Wiley-VCH GmbH
2203153 (2 of 21)
www.advancedsciencenews.com
in 2016.[13] In their work, an external pressure sensor was used
to monitor stress changes during cycling. Subsequently, internal
stress measurements with external pressure sensors have been
reported in various battery materials and systems, including
NCM materials, alloy anode materials, and solid-state electro-
lyte materials. [14] Recently, the fiber Bragg grating (FBG) sen-
sors, which are widely used in liquid LIBs, [15] have been imple-
mented into the solid-state battery system. [16] These in situ or
operando measurements provide a powerful tool for examining
the link between electromechanical eects and battery electro-
chemical performance. Recently, Song et al. performed a com-
prehensive overview of stress sources and the resulting elec-
trochemomechanical eects on composite electrodes and cell
stacks in ASSBs. [17] However, few review articles are available
in the literature to provide a comprehensive summary of the
working principles and experimental devices for stress measure-
ment (such as external pressure sensors and FBG sensors) and
the application of these stress-measuring techniques.
This review examines the recent research progress on
internal stress measurement in sulfide-based ASSBs. First,
volume changes of active materials and solid-state electrolytes
during battery cycling and corresponding stress changes in
electrode level are categorized. Then the basic principles and
measurement devices (including external pressure sensors
and built-in FBG sensors) for internal stress measurement in
sulfide-based ASSBs are introduced. We also summarize and
discuss the relationship between stress changes and electro-
chemical reaction processes in ASSBs. Finally, we present our
views on the future development of internal stress measure-
ment in sulfide-based ASSBs. The overall content of the review
is summarized in a comprehensive schematic diagram in
Figure 1.
2. The Origin of Internal Stress
The electrochemo-mechanical problems caused by internal
stress are frequently ignored in the research of sulfide-based
ASSBs. The internal stress is mainly induced by the coupling of
the volume changes of the active materials during the electro-
chemical process. The volume changes of SSEs caused by Li+
migration are negligible due to their intrinsic ceramic nature.[18]
Therefore, the variation of the active materials will be con-
strained by rigid SSEs, leading to internal stress development.
Figure 1. Comprehensive schematic diagram of electrochemo-mechanical stress origins, stress measuring devices, and applications in sulfide-based
ASSBs.
Adv. Energy Mater. 2022, 2203153
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202203153 by Xiamen University, Wiley Online Library on [27/11/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advenergymat.de
© 2022 Wiley-VCH GmbH
2203153 (3 of 21)
www.advancedsciencenews.com
When active materials with huge volume changes are applied
and the battery capacity is high, the internal stress can even
reach the MPa level.
The volume change of active materials can be described
by the partial molar volume of lithium
()
m
VLi in the corre-
sponding compound.[19] The partial molar volume is defined as
the dierential of the considered homogeneous volume caused
by the addition of lithium (maintaining all other state variables
constant, i.e., pressure p, temperature T, and amount of phase
in the system ni).[14k]
() ()
Li Li
m
V
V
nTpnn
iL
i
(1)
The absolute volume change ΔV of the active materials
during lithium insertion/extraction can be obtained by
integrating.
()
()
∆=
()
()
Li
dL
i
Li
Li
m
1
2
VV n
n
n
(2)
The volume changes derived from crystallographic data for
each active material during electrochemical cycling are summa-
rized in Table 1.
2.1. Cathode Materials
2.1.1. Layered Oxide Materials
LiMO2 (M for transition metals) layered oxides possess a hexag-
onal α-NaFeO2 layered structure with the R-3m space group.[20]
It corresponds to the cubic close-packed (CCP) oxygen arrays
with metal ions occupying octahedral sites, forming consecu-
tive LiO2 and MO2 layers.[21] These layered oxides mainly include
LiCoO2 (LCO), LiNiO2(LNO), LiNixMnyCozO2 (x+ y+ z= 1)
(NMC), and LiNixCoyAlzO2 (x+ y+ z= 1) (NCA), etc. The struc-
tural changes and the corresponding volume changes of these
materials during cycling are discussed in the following section.
LiCoO2 (LCO) is widely used in consumer electronics, and
its theoretical specific capacity is 270 mAh g1. LCO under-
goes complex structural changes during the electrochemical
process.[22] During the Li1-xCoO2 delithiation process, two
hexagonal phases (H1 and H2) and one monoclinic phase
(M1) can be observed. H1 appears in the initial charging stage
(x< 0.04) and is partially converted to H2 when x reaches
0.04, and H2 coexists with H1 until x increases to 0.25.
When x exceeds 0.25, H1 is completely converted to H2, and
when x> 0.47, Li+ is further removed from Li1-xCoO2, which
makes H2 convert to M1. When x exceeds 0.56, M1 trans-
forms into H2 again. When Li-ions are completely extracted
from Li1-xCoO2, CoO2 with a hexagonal monolayer structure
is formed.[23] Among them, the hexagonal hexagonal
transition or the hexagonal monoclinic transition causes
severe volume change. Considering the structural irrevers-
ibility (hexagonal to monoclinic) at the amount of lithium
changes > 0.5mol, its practical capacity is limited to about
145 mAh g1. When charged to Li0.5CoO2, the volume of the
LiCoO2 cell increases by +2%.
LiNiO2 (LNO), an oxide material isomorphic to layered LCO,
is considered a promising cathode material for power LIBs with
high energy density due to its large theoretical specific capacity
(275 mAh g1) and lower cost (avoiding Co).[24] During delithi-
ation/lithiation, LNO undergoes complex structural changes
(hexagonal H1-monoclinic M-hexagonal H2-hexagonal H3
phases) and corresponding huge volume changes (maximum
volume change 9%).[25] Specifically, during the delithiation
process, three two-phase regions: H1-M, M-H2, and H2-H3, are
located at 0.80 x (Li) 0.75, 0.40 x (Li) 0.36 and 0.26 x
(Li) 0.16, respectively. Interestingly, the LNO unit cell exhibits
a nonlinear volume change. Taking the charging process as an
example, the H1-M-H2 phase transition contributes only 3%
of the total 9% volume change. Most of the volume change
(6%) occurs when x 0.26, where the H2–H3 phase transition
results in a 3.8% drop in unit cell volume.[25,26]
By introducing both dopants, Co (to enhance the layered
ordering) and Mn (to stabilize the local structure), into LNO,
Table 1. Volume changes of typical electrode materials during battery
cycling.
Electrode
materials
Reactions Volume change
(%)
Cathode side LiCoO2= 0.5Li + Li0.5CoO21.9[36]
LiNiO2= xLi + Li1xNiO29[25]
Li[Ni1/3Co1/3Mn1/3]O2= xLi + Li1x[Ni1/3Co1/3Mn1/3]O23.4[60]
Li[Ni0.6Co0.2Mn0.2]O2= xLi + Li1x[Ni0.6Co0.2Mn0.2]O25.2[61]
Li[Ni0.8Co0.1Mn0.1]O2= xLi + Li1x[Ni0.8Co0.1Mn0.1]O27.8[61]
Li[Ni0.8Co0.15Al0.05]O2= xLi + Li1x[Ni0.8Co0.15Al0.05]O25.9[62]
LiFePO4= Li + FePO46.5[36]
LiMn2O4= Li + 2MnO27.3[36]
S+2Li = Li2S79[39a]
Se+2Li = Li2Se 98[39a]
Te+2Li = Li2Te 104.7[39a]
CuS+2Li = Cu+Li2S75[39e]
Anode side C6+ Li = LiC613.1[63]
Li4Ti5O12+ Li = Li7Ti5O12 0.2[46]
Si + 4.4Li = Li4.4Si 310[64]
In + Li = InLi 105.6[14j]
Sn +4.4Li = Li4.4Sn 260[64]
Al + Li = LiAl 100[65]
Li foil + xLi = Li1+x4.85 ×
104 cm3 mAh1a)
a)The theoretical volume change of Li foil is calculated as follows.[66]
F
Li
Li
Li
ρ
∆=
VQM (3)
where Q is the lithium deposition capacity (mAh), MLi is the molar mass of
lithium (6.94g mol1), F is the Faraday’s constant, and ρLi is the density of lithium
(0.534g cm3). Therefore, the thickness change of Li foil will reach 48.5 µm at an
areal capacity of 10 mAh cm2.
Adv. Energy Mater. 2022, 2203153
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202203153 by Xiamen University, Wiley Online Library on [27/11/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advenergymat.de
© 2022 Wiley-VCH GmbH
2203153 (4 of 21)
www.advancedsciencenews.com
the developed ternary NMC material has become one of the
most essential layered oxide materials in LIBs technology.[27]
Interestingly, there is an approximately linear relationship
between the unit cell volume change and the Co/(Ni + Co) mole
ratio in NCM materials.[28] The volume change increases as the
Co/(Ni + Co) molar ratio increases (charged to 4.4 V vs Li+/Li
NCN111-1%, NCM811-6%). During delithiation/lithiation,
nickel-rich NCMs (LiNixMnyCozO2: x+ y + z= 1, x 0.8)
undergo similar structural changes to LNO materials.[29] Taking
NCM811 as an example, complex structural changes occur
during the delithiation process: hexagonal to monoclinic
(H1 M), monoclinic to hexagonal (M H2), and hexagon
to hexagon (H2 H3, >4.15 V vs Li+/Li). Among them, the
H2 H3 phase transition leads to dramatic volume change,
resulting in significant volume change/internal stress.[30]
Another ternary layered oxide, NCA, has shown promising
advantages in high-power applications such as electric vehicles
(EVs).[31] Al substitution is considered to improve the thermal
stability of the material due to the stability of Al3+ at the tetra-
hedral sites. During the delithiation process, the NCA cathode
undergoes a series of phase transitions similar to LNO mate-
rials: from the original H1 phase to the monoclinic (M) phase,
followed by the H2 and H3 phases in turn. In particular, for
the Li[Ni0.8Co0.16Al0.04]O2 (NCA80) material, the maximum
volume change during delithiation is about 5.63%, and a dra-
matic volume change is observed at about 4.2V (corresponding
to the H2–H3 phase transition).[32] Similar to NCM materials,
the volume change of NCA materials increases as Ni content
increases (7.10% for NCA88 and 8.37% for NCA95).[32]
2.1.2. Spinel Oxide Materials
Spinel LiMn2O4 has the advantages of low cost and excellent
safety. Its unique 3D lithium-ion transport channel results
in superior rate performance, making it a promising cathode
material for lithium-ion batteries.[33] Severe structural changes
also occur in LiMn2O4 during the Li-ion intercalation and dein-
tercalation process.[34] Three phases (cubic, orthorhombic, and
tetragonal) are observed in the phase transition of LiMn2O4 by
in situ TEM analysis.[35] During the discharging process, the
cubic phase is transformed into the orthorhombic phase in
the first stage and finally into the tetragonal phase. The calcu-
lated volume change from pristine LiMn2O4 to fully delithiated
Mn2O4 is about 7.3%.[36]
2.1.3. Olivine Oxide Materials
Olivine LiFePO4, with a theoretical specific capacity of
170 mAh g1, is widely used in hybrid and electric vehicles due
to its low cost, good cycling performance, and high safety.[37]
LiFePO4 undergoes a two-phase reaction mechanism (the first-
order phase transition of LiFePO4 FePO4 occurs with the
progress of charge/discharge) and a two-phase solid-solution
reaction mechanism (only one phase occurs during the charge
and discharge process) across the flat voltage plateau.[38] Due
to the unique solid solution reaction mechanism, the partial
molar volume of LiFePO4 is constant regardless of the change
in lithium content. The calculated volume change from pristine
LiFePO4 to fully delithiated FePO4 is about 6.5%.[36]
2.1.4. Conversion-Type Materials (Sulfur Based)
Conversion-type materials (sulfur-based materials, such
as sulfur, lithium sulfide, metal sulfides, etc.) are another
common cathode material in sulfide-based ASSBs.[39] These
materials have the advantage of high theoretical specific
capacity, but their operating voltage is lower than that of lay-
ered oxide cathode materials. The “shuttle eect” in liquid
batteries can be eliminated due to the use of solid-state elec-
trolytes, thus the cycling performance of the battery can be
improved. [40] Conversion-type materials exhibit relatively
large volume changes. Taking S as an example, Li2S will gen-
erate during the discharging process. Considering that the
densities of S and Li2S are 2.03 and 1.66g cm3, respectively,
the complete conversion of S to Li2S will result in about 80%
volume expansion.[41]
2.2. Anode Materials
2.2.1. Lithium Intercalation Materials
Graphite materials have the advantages of low cost and low
potential to lithium (0.01–0.2 V vs Li+/Li) and are one of the
most successful anode materials for lithium-ion batteries.[42]
The theoretical specific capacity of graphite is 372 mAh g1.
Multi-step two-phase reactions through various stages (i.e., 1L,
4L, 3L, 2L, 2, and 1) can be observed in the Li-ion intercalation
and deintercalation process, leading to a nonlinear volume
change. The total volume expansion of fully lithiated graphite
(LiC6) is 13.2%.[43]
Li4Ti5O12 has good rate performance and high safety and is
often used in electric buses and other fields.[44] The theoretical
capacity of LTO is 175 mAh g1, and a two-phase reaction of
Li4Ti(IV)5O12Li7Ti2(IV)Ti3(III)O12 occurs during cycling.[45] The
volume change of LTO during cycling is only 0.2%, and it is
called a zero-strain material.[46] The zero-strain nature of LTO
allows it to be used as a counter electrode for pressure meas-
urement analysis of working electrodes.[14b,f,47]
2.2.2. Lithium-Alloying Materials
Si has an extremely high theoretical capacity of 4200 mAh g1,
corresponding to Li22Si5. However, the fully lithiated Si phase at
room temperature is Li15Si4, corresponding to 3579 mAh g1.[48]
Crystalline silicon is amorphized during the first lithiation pro-
cess and forms amorphous silicon (LixSi) through a two-phase
reaction that appears around 170 mV versus Li+/Li. Further
lithiation leads to the crystallization of amorphous silicon into
Li15S4 below about 60mV versus Li+/Li.[49] The volume expan-
sion of Si during the first alloying process is nearly 300%.[50]
Si undergoes irreversible structural changes during the first
charging-discharging process, e.g., during the first charging
process, Si undergoes a crystalline to amorphous transition at
Adv. Energy Mater. 2022, 2203153
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202203153 by Xiamen University, Wiley Online Library on [27/11/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advenergymat.de
© 2022 Wiley-VCH GmbH
2203153 (5 of 21)
www.advancedsciencenews.com
a lower electrode potential than in subsequent cycles, during
which the material remains amorphous.[51]
Indium anode is also a widely used alloyed material in
sulfide-based ASSBs. Indium can form a lithium-indium alloy
during cycling and has a stable potential (0.62V vs Li+/Li).[52]
Besides, the lithium-indium alloy also has good mechanical
properties and chemical stability with sulfide SSEs.[53] Indium
is lithiated during the charging process into a slightly nonstoi-
chiometric InLi1-x phase, accompanied by a transition from the
pure indium tetragonal phase to the cubic InLi1x phase. The
volume of the tetragonal unit cell of InLi is 113.23 Å3, which
is approximately two times larger than that of the pure In
tetragonal phase (55.069 Å3). Therefore, the change from In
(m
V
= 15.71 cm3 mol1) to InLi ( m
V
= 23.597 cm3 mol1 of the
stoichiometric phase) is accompanied by a volume expansion
of 105.6%.[14j]
Lithium metal is considered an ideal anode material because
of its high theoretical specific capacity (3860 mAh g1) and the
lowest redox potential (3.04V vs SHE).[54] However, Li metal
anodes suer from large volume changes (related to the amount
of deposition) during cycling. The infinite volume change
of pure lithium metal stems from its “hostless” nature.[55] In
a constrained sulfide-based all-solid-state lithium metal bat-
tery, the stress evolution caused by this large volume change
of the Li metal anode during cycling is considered to be the
main cause of the nonuniform distribution of current density
of the electrodes and results in the formation of lithium den-
drites (micro-short circuits) as well as the mechanical failure of
the battery. However, the complex stress evolution of Li metal
anodes, including Li-alloy anodes and composite Li electrodes
with flexible Li hosts during electrochemical cycling, is not yet
well understood and needs careful examination in the future.
2.3. Solid-State Electrolyte Materials
Due to the narrow electrochemical stability window, the sulfide
SSEs without any modification, will inevitably undergo decom-
position while mixing with the conductive carbon materials in
the layered oxide composite cathode, resulting in some irrevers-
ible stress changes.[7b,c,56] For example, during charging, sulfide
SSEs will undergo oxidation reactions accompanied by volume
reduction.[14f ] In layered oxide composite cathode, the volume
change caused by the oxidation of the sulfide SSEs is mar-
ginal for the total internal stress change compared with that
of the active materials.[14f ] However, the dierences in volume
changes and mechanical parameters (such as Young’s modulus
and bulk modulus, etc.) between the SSEs and the layered oxide
will cause uneven local stress distribution inside the composite
cathode, resulting in contact loss at the three-phase interface.[17]
It is worth noting that the operating voltage range of ASSBs
is relatively narrow when sulfur-based conversion-type active
materials are used in the composite cathodes (such as 1.5–3V
for sulfide-based all-solid-state Li–S batteries).[57] In this case,
reversible redox reactions of sulfide SSEs may occur, providing
a considerable reversible capacity.[58] Therefore, the contribu-
tion of the stress change caused by the reversible redox of the
sulfide SSEs to the total electrode stress change is no longer
negligible.
On the anode side, the electrochemical reduction will occur
when the voltage is as low as the electrochemical decomposi-
tion voltage of sulfide SSEs. For example, in a composite anode
consisting of LGPS, when voltage is below 0.6V versus Li/Li+,
LGPS will be electrochemically reduced to Li3P, LixGe, and
other products, resulting in an increase in internal stress.[59] In
particular, a study reported by Jung et al. showed that in the
LGPS/Gr composite anode, the internal stress change caused
by the electrochemical reduction of LGPS was comparable to
that caused by Li+ intercalation into graphite. [14c] The measure-
ment of the stress change caused by the volume changes of the
SSEs and the corresponding electrochemo-mechanical issues
will be discussed in detail in Section4.3.
3. Principles and Devices of Internal Stress
Measurement
3.1. Principles of Internal Stress Measurement
As mentioned above, when a volume-constrained all-solid-
state battery is cycling, the internal stress of the battery system
changes due to the volume change of the cathode and anode
active materials. The volume changes of the active materials
and the pressure change of the battery system can be bridged
by the mechanical properties of battery components: Young’s
modulus (E), shear modulus (G), bulk modulus (K), and Pois-
son’s ratio (υ).[67] The mechanical parameters of major elec-
trode materials and sulfide SSEs are listed in Table 2. It is
worth noting that the elastic properties of electrode materials
are greatly aected by the degree of lithiation, which makes it
challenging to evaluate real-time stress changes in ASSBs.[63,68]
Janek etal. proposed a simple mechanical model to estimate
the relationship between material volume change and pressure
change in ASSBs.[14k] They assumed that the all-solid-state bat-
tery is made of one single material (β-Li3PS4 glass-ceramic),
and the solid-state electrolyte is elastic enough to behave like
a fluid in a system, which is constrained on all sides. The pres-
sure change inside the battery can then be deduced by using
the isothermal compressibility factor κT consisting of Poisson’s
ratio υ and Young’s modulus E.
κ
υ
()
== 1312
T
KE
(4)
By integrating the isothermal compressibility, Hencky’s loga-
rithmic strain can be obtained. Where V0 is the initial volume,
and εvol =ΔV/V0 is the volumetric strain.
∫∫
κ
=−
dd
T
00
V
Vp
V
V
p
p
(5)
κεκ
−+
=− +ln 11ln 11
0T
vol
T
pV
V (6)
The approximate relationship between the internal stress
change, the bulk strain εvol, and the bulk modulus K of SE can
be obtained by linearizing the logarithmic strain.
Adv. Energy Mater. 2022, 2203153
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202203153 by Xiamen University, Wiley Online Library on [27/11/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advenergymat.de
© 2022 Wiley-VCH GmbH
2203153 (6 of 21)
www.advancedsciencenews.com
κεκε
∆≈
=− ∆=
11
or
0T
vol
T
vol
p
V
V
pK
(7)
In an actual experiment, the internal stress change of the
system can be measured by external pressure sensors or built-
in FBG sensors.
3.2. Stress Measuring Devices
Until now, the stress measuring devices reported in sulfide-based
ASSBs mainly fall into two categories: the external pressure
sensor type and the built-in optical FBG sensor type. The former
detects the stress change during cycling by placing an external
pressure sensor outside the battery system, while the latter uses
a built-in optical FBG sensor to measure the battery system pres-
sure indirectly through the conversion between the Bragg wave-
length and the physical parameters of the FBG sensor.
3.2.1. External Pressure Sensor
Some pressure-measuring devices in sulfide-based ASSBs
reported so far are shown in Figure 2a,b.[14a,b] A typical pressure-
measuring device reported by McDowell et al. consists mainly
of three parts: a volume-constrained mechanical framework, a
battery cell, and a pressure sensor (Figure2a).[14b] Among them,
in order to ensure the reliability of the stress measurement,
a reasonably designed mechanical framework is reasonably
designed. This fixture can apply uniform pressure to the battery
by tightening the four bolts with a precision torque wrench.
Therefore, the influence caused by the relaxation of the battery
components can be minimized.
At present, commercially available pressure sensors mainly
include capacitive sensors,[82] resistive sensors,[83] piezoelectric
sensors,[84] and piezoresistive sensors.[85] These sensors can
convert the measured pressure into a corresponding electrical
signal output (voltage or current).
Take the most widely used force-sensing resistor (FSR) pres-
sure sensor as an example. The metal strain gauge will deform or
stretch under external pressure, changing its resistance value. The
relationship between the deformation of the metal strain gauge
and the change in resistance can be expressed by Equation8:[86]
=
R
R
k
L
L
(8)
where k represents the sensitivity coecient; R represents the
resistance of the metal strain gauge; ΔR represents the change
in resistance; L represents the length of the strain gauge; ΔL
represents the change in the length of the strain gauge.
Table 2. Mechanical parameters of major electrode materials and sulfide SSEs.
Materials Young’s modulus
E [GPa]
Poisson’s ratio
[ν]
Shear modulus
G [GPa]
Bulk modulus
K [GPa]
Cathode materials LiCoO2191.0 0.24 80 122.4[69]
LiNi0.33Co0.33Mn0.33O2199.0 0.25 78 132.6[70]
LiFePO4117.8 0.30 45.5 98.2[71]
LiMn2O4194.0 0.26 77.0 134.7[34c]
α-S812.8 0.23 5.2 8[72]
Li2S 76.62 0.18 32.35 39.9[73]
TiS2 (1T)a) 228 0.11 102 97.4[74]
MoS2 (2H)a) 230 0.3 88 191.7[74]
Sulfide SSEs β-Li3PS429.5 0.29 11.4 23.3[75]
75Li2S–25P2S523 0.32 8.7 21[76]
Li7P3S11 21.9 0.35 8.1 23.9[75]
Li10GeP2S12 21.7 0.37 7.9 27.3[75]
Li6PS5Cl 22.1 0.37 8.1 28.7[75]
Li6PS5Br 25.3 0.35 9.3 29.0[75]
Li6PS5I 30.3 0.33 11.3 29.9[75]
Anode materials Li 4.9 0.42 4.2 11.0[77]
In 12.6 0.45 4.4 42.2[78]
Si 96.0 0.29 62.0 90.0[79]
Sn 50.5 0.34 21.1 54.4[80]
Li13Sn548.7 0.21 24.4 33.3[80]
Li2.25Al 62.4 0.21 [34c]
LTO 181.0 0.25 73.1 125.1[81]
Graphite 32.0 0.31 12.0 28.1[63]
a)2H and 1T represent trigonal prismatic coordination and octahedral coordination configuration, respectively.
Adv. Energy Mater. 2022, 2203153
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202203153 by Xiamen University, Wiley Online Library on [27/11/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advenergymat.de
© 2022 Wiley-VCH GmbH
2203153 (7 of 21)
www.advancedsciencenews.com
The key specifications of pressure sensor performance
include linearity, hysteresis, temperature sensitivity, creep and
repeatability,[87] etc.
1. Linearity: Linearity is defined as the parameter of the accu-
racy of the corresponding relationship between the load and
the electrical signal output of the sensor.
2. Hysteresis: Hysteresis can be determined by observing the
output signal of the sensor when it is loaded and unloaded.
Hysteresis will significantly aect the symmetry of the meas-
ured pressure.
3. Temperature sensitivity: Temperature sensitivity is defined
as the stability of output sensitivity when the ambient
temperature changes. It is generally measured in the unit
Figure 2. a) Schematic diagram of the assembly and pressure measurement of an anode/LPSC/cathode cell. Reproduced with permission.[14b] Copyright
2021, Elsevier. b) Schematic of in situ pressure-monitoring in an ASSB. Reproduced with permission.[14a] Copyright 2022, Elsevier. c) Image of a fully
assembled lithium-ion battery coin cell with FBG sensor integrated. Reproduced according to the terms of the CC-BY license.[109] Copyright 2017, The
Authors, published by MDPI AG. d) Photographs of the optical fiber embedded within the InLix anode of an InLi|Li3PS4|LTO ASSB. Reproduced with
permission.[16] Copyright 2022, The Authors, published by Springer Nature. e) The working principle of an FBG optical sensor. Reproduced with permis-
sion.[16] Copyright 2022, The Authors, published by Springer Nature. f) The scheme of the birefringence phenomena in an FBG sensor when a relatively
high external pressure is applied. Reproduced according to the terms of the CC-BY license.[16] Copyright 2022, The Authors, published by Springer Nature.
Adv. Energy Mater. 2022, 2203153
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202203153 by Xiamen University, Wiley Online Library on [27/11/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advenergymat.de
© 2022 Wiley-VCH GmbH
2203153 (8 of 21)
www.advancedsciencenews.com
of drift generated within the range of every 10 °C, such
as 0.05% F.S./10 °C.Temperature sensitivity drift can be
reduced by temperature compensation.
4. Creep and repeatability: Creep is related to the time-depend-
ent permanent deformation of a material when subjected to
constant pressure. Repeatability refers to producing reliable
results even over a long measuring time.
Therefore, selecting an appropriate pressure sensor
according to the dierent solid-state battery systems is cru-
cial. Nevertheless, the performance parameters of pressure
sensors are rarely introduced in the reported literature. In
addition, another factor aecting pressure measurement
accuracy is pressure baseline changes caused by mechanical
relaxation.[14m] Some work compensated for baseline changes
in the measured pressure curve,[14j,k] while others rested the
battery system for a long time enough to make the baseline
changes negligible. [14b]
3.2.2. Built-in Fiber Optic Sensor
As shown in Figure2c,d, the built-in optical FBG sensor stress
measurement device needs to be skillfully implanted into the
battery system. As illustrated in Figure2e, when light travels
through the optical fiber, the FBG sensor acts as a reflector for
Bragg wavelength (λB), which is defined as λB = 2neΛ, where
ne is the eective refractive index and Λ is the Bragg grating
period. Any changes in temperature (T), hydraulic pressure (P),
or strain (ε) that occur around the FBG sensor will modify ne
and/or Λ, which will be translated into a change in reflected
wavelength, visualized as a peak shift (ΔλB).[15a,c,d,f,g,88]
FBG sensors can obtain rich thermal/mechanical informa-
tion.[89] However, when we only want to focus on certain system
information, the richness of information will be a double-edged
sword. Therefore, when we want to detect stress changes in solid-
state batteries with FBG sensors, some settings must be made to
ignore the contributions coming from T and hydraulic P varia-
tions. [16] First, the battery tests should be performed in a temper-
ature-controlled climate chamber to eliminate any fluctuations
in ambient temperature and ensure that the temperature change
of the battery during cycling is negligible. Besides, low-loading
active materials should be used to minimize inherent heat
release and gas generation to ensure limited hydraulic pressure.
When the FBG sensor is strained under these conditions,
the basic principle of the FBG sensor can be expressed by
Equation9:[16]
λ
λρε υε
()
()
=− =− −+
11 2
B
B,0
e
eff
2
12 11 12
np pp (9)
where λB,0 is the Bragg wavelength at the initial moment, ρe is the
eective photo-elastic coecient, p11 and p12 are the strain optical
coecients of the fiber, and ν is the Poisson’s ratio. The above
parameters are known for silica fibers. The measured strain ε can
be directly converted to stress by using Hooke’s law: σ= εE, where
E is Young’s modulus of the silica fiber, equal to 69.9GPa. In addi-
tion, the stress-wavelength shift curve of FBG sensors can be cali-
brated first to obtain the real-time stress value in the linear region.
Interestingly, peak splitting occurs in FBG sensors when a
relatively high external pressure is applied. This phenomenon,
known as birefringence, is caused by the elliptical deformation
of the fiber driven by an external force field (Figure 2f). This
symmetry break causes the initial eective refractive index to be
anisotropic, resulting in two distinct refractive index values (nx
and ny), where the x polarization (nx) varies much more than the
y polarization (ny). Thus, when a suciently high lateral load is
applied to the FBG sensor, the single resonant peak of the FBG
(λB, given by λB = 2neΛ) splits into two peaks (λx and λy).
The birefringence (B) is given by Equation10: [16]
==+
∆−
eff,0
0
yx
eff,0
B
nn
nB
nn
n (10)
where n|| and n are the refractive index in parallel and per-
pendicular directions to the applied load, Δnx and Δny are the
refractive index changes in x and y light polarization due to the
externally applied load, and ne,0 is the initial eective refrac-
tive index. B0 is the birefringence caused by grating fabrication
and is negligible for low birefringence FBG sensors. By decou-
pling λx and λy and combining them with a calibrated pressure-
wavelength working curve, the local stress information of the
electrode can be obtained.
Table 3 compares the advantages and disadvantages of the
external pressure sensors and built-in FBG sensors. In gen-
eral, the main advantage of the external pressure sensor is that
it is simple to operate and can directly and nondestructively
track and measure the stress change in the battery system.
However, the stress change measured by the external pres-
sure sensor is only in the axial direction of the cell scale. Since
the built-in FBG sensor needs to introduce the optical fiber
into the battery system, the operation is complicated, and the
battery structure will inevitably be damaged. The biggest fea-
ture of the FBG sensor is that the local stress in the axial and
longitudinal directions can be obtained at the electrode scale,
which is currently impossible for external pressure sensors.
4. Application
Accurately measuring internal stress changes during battery
cycling is crucial to understanding the electrochemo-mechan-
Table 3. Advantages and disadvantages of the external pressure sensors and built-in FBG sensors.
Pressure-measuring devices Destructive Operational complexity Stress scale Local stress
External pressure sensors × × Device level ×
Build-in FBG sensors Material level (Axial + longitudinal)
Adv. Energy Mater. 2022, 2203153
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202203153 by Xiamen University, Wiley Online Library on [27/11/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advenergymat.de
© 2022 Wiley-VCH GmbH
2203153 (9 of 21)
www.advancedsciencenews.com
ical eects. Depending on the degree of volume change caused
by the reaction mechanism (intercalation/conversion/alloying
reactions) and the variation in mechanical property parame-
ters, the internal stress changes from dierent active materials
into devices tend to vary widely. The magnitude of the change
in internal stress will aect the degree of electrode structure/
interface evolution, resulting in mechanical degradation of
battery capacity. In addition, the mechanical information
provided by the pressure measurement is also related to the
electrochemical behaviors, such as the reaction mechanism of
the active materials, the state of charge (SoC), and the state of
health (SoH) of the battery. Therefore, the pressure measure-
ment can bridge the gap between mechanics and electrochem-
istry. In the following sections, the application of pressure
measurement to sulfide-based ASSB materials and devices is
reviewed.
4.1. Cathode Active Materials: Electrochemistry and Mechanics
Layered oxide cathode materials are widely used in sulfide-
based ASSBs. Janek’s group systematically measured the
pressure changes of different materials (NCM, NCA, and
LCO) during cycling by using a zero-strain LTO anode as the
counter electrode.[14k] As shown in Figure 3a–c, the experi-
mentally measured pressure changes for these three mate-
rials are comparable (102 MPa level, depending on areal
capacity). In addition, the changing trend of measured pres-
sure of the three materials is consistent with the volume
change obtained by crystallography; that is, the measured
pressure of NCM and NCA materials decreases during the
charging process and increases during the discharging
process, while the LCO material shows the opposite trend.
When matched with the same anode, the opposite pressure
changes of NCM and LCO materials will lead to different
electrochemo-mechanical effects, resulting in different
degrees of mechanical damage to the battery structure.
Interestingly, McDowell’s group recently reported that when
lithium metal is used as an anode, the ASSB with NCM/LCO
as the active material shows completely different micro-
short-circuit behavior.[90] Under identical operating condi-
tions (current density of 2 mA cm2 and areal capacity of
6 mAh cm2), the Li/LPSC/NMC811 battery exhibited
micro-short-circuit behavior in the first charging process
(indicated by a sudden voltage drop and an overcharged
capacity far exceeding the theoretical areal capacity). In
contrast, the Li/LPSC/LCO battery was charged normally
in the first cycle. This phenomenon is very interesting as
it may incorporate information on the micro-short-circuit
failure mechanism caused by the “internal stress” (stress
coupling between cathode and anode) generated in Li metal
full cells, which cannot be obtained in the Li/Li symmetric
cell.[16] The structure of the symmetrical cell is destined to
be incapable of considering the influence of internal stress
(one side plating and the other side plating, the system is
in an internal stress balance). Recently, Budiman et al.
performed electrochemo-mechanical analysis on NCM/
LGPS/LiIn and LCO/LGPS/LiIn configuration batteries by
finite element analysis (FEA).[91] In this study, the authors
simulated the stress distribution within the SSE caused
by electrode volume changes and further mechanical
damage behavior, including plastic deformation, interfacial
debonding, and crack propagation. The simulation showed
that the most severe mechanical damage to SSE occurred
when cathode and anode active materials exhibited oppo-
site volume changes (such as the NCM cathode shrinkage
and the LiIn anode expansion during the charging pro-
cess). While different electrochemo-mechanical effects in
NCM/LCO-Li full cells make intuitive sense when LiIn is
replaced by a Li metal anode, their influence on the lithium
plating/stripping behavior or SSE mechanical damage and
the resulting micro-short-circuit of the Li metal anode still
needs to be further studied. Although challenging, tracking
the pressure changes in the battery during cycling may
shed light on understanding these complex electrochemo-
mechanical effects. The crystal structure design of Ni-rich
layered cathode materials greatly influences their elec-
trochemo-mechanical effects. Jung et al. investigated the
structure evolution of two different Ni-rich materials, a
commercial-grade Li[Ni0.80Co0.16Al0.04]O2 (NCA80) having
randomly oriented grains and a full-concentration gradient
Li[Ni0.75Co0.10Mn0.15]O2 (FCG75) having radially oriented
rod-shaped grains by in-situ pressure measurement.[47] As
shown in Figure3d, NCA80 exhibited a lower pressure drop
(-0.15MPa) than FCG75 (-0.30MPa) after the first charging
process. This result directly indicates the formation of voids
or cracks in NCA80. The pressure of NCA80 is higher than
that of FCG75 after one cycle, although fewer Li+ ions are
intercalated in the lattice of NCA80 compared to FCG75
(corresponding to a lower discharge capacity), suggesting
severe mechanical degradation of the NCA80 electrode.
Compared with layered oxide materials, the volume
changes of conversion cathode materials are much larger.
Janek’s group monitored the pressure changes during
cycling in a Li0.3In0.7/Li7P3S11/S: Li7P3S11: Ketjen black
50:30:20 wt% cell (Figure 3e).[13] In their experiment, a
pressure decrease during discharging was observed. Inter-
estingly, the authors seem perplexed by this result, as
they expected the S8 Li2S reaction to being a volume-
increasing process. In fact, since the pressure sensor moni-
tors the resultant force in the direction of σ11, the decrease
in net pressure during the discharging process means that
the net volume change of the battery is dominated by the
Li–In anode. When a conversion-type material with a larger
volume change (e.g., Se, which exhibits a 98% volume
change when fully lithiated) is used on the cathode side,
the net pressure change is likely to be controlled by the
cathode. Recently, Sun’s group confirmed this possibility
in an all-solid-state Li–Se battery (Figure 3f).[14h] In their
experiments, the net pressure change of the battery was
determined by the volume change of the Se cathode. How-
ever, since most of the conversion active materials do not
contain lithium, zero-strain LTO cannot be directly used as
the counter electrode, and it is still a challenge to measure
the pressure change during the cycling of a single conver-
sion active material electrode.
Adv. Energy Mater. 2022, 2203153
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202203153 by Xiamen University, Wiley Online Library on [27/11/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advenergymat.de
© 2022 Wiley-VCH GmbH
2203153 (10 of 21)
www.advancedsciencenews.com
Some sulfide SSEs can undergo reversible redox reactions
within a specific voltage range and have also been reported as
conversion-type cathode active materials. However, the complex
redox reaction mechanism remains unclear. Based on the high
sensitivity and fast response between electrochemical reactions
and specific volume changes (pressure changes), tracking elec-
trochemo-mechanical changes via in situ pressure measure-
ment can be used to reveal possible reaction pathways. Wang
et al. reported an ASSB (In/InLi|LPSCB|LPSCB-MWCNTs)
using Li6PS5Cl0.5Br0.5 as the cathode active material.[14i] In
this work, pressure measurement is first used to reveal the
reaction mechanism. As shown in Figure 3g (bottom), the
measured pressure curves show nonlinear behavior with
three distinct regions during the charging process. Consid-
ering that the partial molar volume change of the Li–In anode
(In/InLi
)
rm
V=+7.89 cm3 mol1 is constant, it is suggested
Figure 3. a–c) Change of the net stress σ11 for ASSBs with dierent electrode materials. Reproduced with permission.[14k] Copyright 2018, The Royal Society
of Chemistry. d) Operando electrochemical pressiometry profiles for FCG75 and NCA80 during first charge-discharge at 0.1 C and 30 °C. Reproduced
with permission.[47] Copyright 2019, Wiley-VCH. e) Cycling behavior and pressure change of a Li0.3In0.7/Li7P3S11/S:Li7P3S11:Ketjen black (50:30:20 wt%)
ASSB cycled at room temperature. Reproduced with permission.[13] Copyright 2016, American Chemical Society. f ) Discharge/charge profiles combined
with the in-situ cell pressure evolution of an all-solid-state Li–Se battery. Reproduced with permission.[14h] Copyright 2022, Wiley-VCH. g) Exemplary
potential profile combined with the in situ measured pressure change of an In/InLi|LPSCB|LPSCB-MWCNTs cell during cycling (top) and a magnification
of the data for one cycle (bottom). Reproduced according to the terms of the CC-BY license.[14i] Copyright 2021, The Authors, published by Wiley-VCH.
Adv. Energy Mater. 2022, 2203153
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202203153 by Xiamen University, Wiley Online Library on [27/11/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advenergymat.de
© 2022 Wiley-VCH GmbH
2203153 (11 of 21)
www.advancedsciencenews.com
that at least three reactions with dierent volume changes are
involved on the cathode side. The potential reaction mechanism
can be further given by comparing the partial molar volume
change of the cathode with the partial molar volume change
of the Li–In anode. If we suppose that the PS43 phase formed
by a reaction of Li2S and P0 (representing as (Li2S)0.75(P2S5)0.25)
should first be converted to (Li2S)0.67(P2S5)0.33 (representing as
Li4P2S7, P2S74), and finally to (Li2S)0.50(P2S5)0.50 (representing as
Li2P2S6, P2S62) during charging. Therefore, the corresponding
reaction volume changes can be calculated as follows:
Reaction (1)-Li3PS4 (LPS) phase formation:
()
()
∆+→+
+
=−
+−
4LiS 2LiS PS 5Li5
5.60 cm mol
rm 22
0.75 25
0.25
31
VP e (11)
Reaction (2)-LPS oxidation I:
() ()
() ()
∆→
++ +
=−
+−
4LiS PS 3LiS PS S2Li 2
8.50 cm mol
rm 20.75 25
0.25 20.67 25
0.33
31
Ve
(12)
Reaction (3)-LPS oxidation II:
() ()
() ()
∆→
++ +
=
+−
3LiS PS 2LiS PS S2Li 2
5.76 cm mol
rm 20.67 25
0.33 20.50 25
0.50
31
Ve
(13)
Only reaction (2) can lead to a decrease in Δp upon charging.
Therefore, here, we assumed that the oxidation of the LPS glass
phase starts in region 2. Consequently, the formation of the
LPS glassy phase occurs in region 1. Although the complete
electrochemical reaction pathway is more complex, the above-
simplified reaction volume can explain the observed Δp trend,
providing a new perspective for revealing electrochemical reac-
tion mechanisms and pathways.
4.2. Anode Active Materials: Electrochemistry and Mechanics
Lithium metal is considered the “Holy Grail” among anode
materials. Due to the thermodynamic instability between the
sulfide SSEs and Li metal, however, interphases will be gener-
ated at the interface.[92] Due to the partial molar volume dier-
ence between the interphase and the Li metal, the formation
and evolution of the interphase at the interface will aect the
battery pressure. Therefore, pressure measurement can pro-
vide a new electrochemo-mechanical perspective on interface
evolution. Lee etal. systematically investigated two types of rep-
resentative SSEs (LSPS and LPSCl)/Li interfaces evolution by
in situ monitoring the pressure changes in SSE/Li symmetric
cells (Figure 4a).[14m] Compared with LPSCl symmetric cells,
LSPS symmetric cells exhibit a more obvious pressure drop
at open circuit voltage (OCV). This is owing to the formation
of electrically conductive Li–Sn alloys upon contact between
LSPS and lithium metal, which allows the interfacial reactions
to continue. The authors further investigated the relationship
between pressure changes and the electrochemical processes
in LSPS symmetric cells. Due to the formation of the electro-
chemical interface, as shown in Figure4b, the symmetric cell
pressure drops more rapidly. Interestingly, when the short cir-
cuit caused by the penetration of lithium dendrites occurs in
the symmetric cell, the pressure change tends to plateau (dP/dt
approaches zero). This electrochemomechanical response fea-
ture suggests that pressure measurement can serve as a pow-
erful tool to detect short-circuit failures in Li-metal ASSBs.
A recent work reported by Lim et al. confirms this possi-
bility. As shown in Figure 4c,d, the authors monitored the in
situ pressure changes during cycling in both NCM/uncoated
Li and NCM/LixIn-coated Li batteries.[93] Since the volume
change during Li plating/stripping is much larger than that of
the NCM materials, the periodic linear pressure changes are
mainly dominated by the Li metal anode when the battery is
cycled normally. In NCM/uncoated Li battery, the pressure no
longer changes linearly when a short circuit behavior occurs.
This is caused by the penetration and growth of Li dendrites
in the mechanical stress-caused defects within the SSE. The
authors further normalized the pressure change dierences
(Δ(ΔP)) with respect to the discharge capacity (Q) at each cycle
(denoted as Δ(ΔP)Q) (Figure4e). The capacity-normalized pres-
sure change dierence Δ(ΔPQ) was calculated by the following
equation: Δ(ΔPQ) =Δ(ΔP)/Qdischarge where Qdischarge and Δ(ΔP) is
corresponding discharge capacities and pressure change dier-
ence at a corresponding cycle. The decrease in Δ(ΔP)Q is associ-
ated with the penetrating growth of Li dendrites within SSE.
Therefore, the growth mechanism of Li dendrites can be eec-
tively predicted by monitoring the Δ(ΔP)Q value.
Due to the severe interfacial reactions and microshort cir-
cuit issues, lithium metal is dicult to be practically applied in
the short term. As an alternative, Li-alloy anodes are expected
to achieve high energy density ASSBs due to their high spe-
cific capacity and low potential. Among them, Li–In alloys are
the most widely used alloy anode materials in the laboratory.
Although Li–In anode and sulfide SSE are thermodynami-
cally and kinetically stable, Li–In anode also exhibits huge
volume changes during cycling, causing serious electrochemo-
mechanical problems in batteries. Recently, Tarascon’s group
introduced the FBG sensors into ASSBs and systematically
measured the stress evolution inside the Li–In anode and at the
anode/SSE interface (with LTO as the counter electrode).[16] As
a comparison, the author also used an external pressure sensor
to measure the pressure change. In the pressure measure-
ment inside the Li–In anode, although the pressure changes
measured by the external pressure sensor and the FBG sensor
are symmetrical and periodic, the stress change monitored
in the FBG sensor is an order of magnitude larger than that
of the external pressure sensor (Figure4f,g). It is particularly
worth pointing out that in this work, for the first time, the
local stress evolution between the Li–In anode/SSE interface
was monitored through the “birefringence eect” of the FBG
Sensor (Figure 4h,i). Measuring such local stress evolution
at the interface is essential for a deeper understanding of the
electrochemo-mechanical eects in ASSBs. In addition to Li–In
anode, Han etal. systematically investigated the pressure evolu-
tion in ASSBs with various composite alloy anodes (Sb, Sn, and
Si) (Figure4j–l).[14b] Since the volume change of NCM111 on the
cathode side is smaller than that of the alloy anode, the net pres-
sure change is controlled by the alloy anode. A typical feature of
Adv. Energy Mater. 2022, 2203153
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202203153 by Xiamen University, Wiley Online Library on [27/11/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advenergymat.de
© 2022 Wiley-VCH GmbH
2203153 (12 of 21)
www.advancedsciencenews.com
alloy anodes is the structural change that occurs during the first
cycle, which is reflected in the mechanical information as the
most severe stress hysteresis and irreversibility during the first
cycle. This result indicates the sensitivity of the pressure meas-
urement to the response of electrode structure evolution. It is
worth noting that the particle size for the stress evolution of
the Si anode was studied intensively in this work. The authors
found that the Si nanoparticles showed less stress change than
the Si microparticles by normalizing the stress change and the
areal capacity. Smaller normalized stress variation may lead
to less electrochemo-mechanical degradation, thus improving
cycling stability.
Figure 4. a) Stack pressure evolution after pressing to an initial stack pressure of 30MPa and holding at open circuit for two dierent symmetric cells
based on LSPS (red) and LPSC (blue). Reproduced with permission.[14m] Copyright 2021, American Chemical Society. b) Voltage curves (red) and stack pres-
sure profiles (blue) of two LSPS symmetric cells (operated at 0.5mA cm2 (solid lines), and held at OCV (dotted lines). Reproduced with permission.[14m]
Copyright 2021, American Chemical Society. c,d) Charge-discharge voltage profiles NCM/Li ASSB and corresponding pressure changes for c) uncoated and
d) In/LixIn-coated thin LMAs. Reproduced with permission.[93] Copyright 2022, Elsevier. e) Capacity-normalized pressure change dierence (Δ(ΔP)Q) as a
function of cycle number. Reproduced with permission.[93] Copyright 2022, Elsevier. f) Time-resolved voltage (top), external cycling pressure (middle), and
an internal optical signal (bottom) for an LTO-LiIn ASSB cycled at C/30 (5.83mA g1) and 25 °C in an operando mode. Reproduced with permission.[16]
Copyright 2022, The Authors, published by Springer Nature. g) Stress evolution obtained from the FBG sensor placed inside Li–In anode by demodulating
the optical signal (blue curve). Reproduced according to the terms of the CC-BY license.[16] Copyright 2022, The Authors, published by Springer Nature.
h) 2D stack view of the operando collected spectra by the FBG sensor placed between LTO/Li–In interface, with the corresponding galvanostatic charge/
discharge cycle. A typical birefringence phenomenon could be found. Reproduced according to the terms of the CC-BY license.[16] Copyright 2022, The
Authors, published by Springer Nature. i) Galvanostatic cycle (top), λx and λy evolution (middle), and operando stress evolution obtained internally by
the FBG sensor and with the experimental calibration of the sensor (bottom). Reproduced according to the terms of the CC-BY license.[16] Copyright 2022,
The Authors, published by Springer Nature. j–l) Galvanostatic cycling of j) Sb, k) Sn, and l) Si composite anodes in ASSBs with LPSC solid electrolyte and
NMC-111 cathodes, along with the measured uniaxial stress (stack pressure) changes in each cell. Reproduced with permission.[14b] Copyright 2021, Elsevier.
Adv. Energy Mater. 2022, 2203153
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202203153 by Xiamen University, Wiley Online Library on [27/11/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advenergymat.de
© 2022 Wiley-VCH GmbH
2203153 (13 of 21)
www.advancedsciencenews.com
4.3. Inactive Components: Electrochemistry and Mechanics
Besides the active materials, the mechanical properties of
the inactive components (SSE and binder) in the composite
electrode are equally critical to the electrode structure stability.
Wang etal. investigated the eect of thiophosphate SSE with
dierent crystallinity (7525-glass and crystalline β-Li3PS4) on
the electrochemical performance of ASSBs.[14g] As illustrated in
Figure 5a, the redox processes of two SSEs can be correlated
with pressure changes by cyclic voltammetry (CV) test and in-
situ pressure monitoring. β-Li3PS4 has a more severe redox than
7525-glass, which will lead to a more severe volume change
(pressure change). Due to the mismatch in the volume change
between the SSE and the cathode active material, more severe
physical contact loss and cracks will occur (Figure5b), deterio-
rating the electrochemical performance of the battery. Similar
conclusions were obtained in work reported by Teo et al. The
authors investigated the electrochemo-mechanical behavior of
glassy (1.5Li2S–0.5P2S5–LiI) and crystalline (Li6PS5Cl) thiophos-
phate SSEs by in situ pressure measurement[94] (Figure 5c).
Likewise, crystalline Li6PS5Cl has more pronounced redox
reactions and pressure changes. Compared with sulfide SSEs,
halide SSEs have recently attracted great attention due to their
wider electrochemical stability window and better chemical
stability. However, the dierence in mechanical properties
between sulfide and halide SSEs remains unclear. Han et al.
investigated the electrochemo-mechanical eects of sulfide
(Li6PS5Cl0.5Br0.5, LPSX) and halide (Li3YCl6, LYC) SSEs on bat-
tery performance by pressure measurement.[14f ] As shown in
Figure5d, the poly-NCA electrode with LPSX showed an abrupt
drop in pressure at the beginning of the first charging process.
This feature is more pronounced in the capacity-derivative dif-
ferential pressure (or dierential electrochemical pressiom-
etry (DEP)) curve. In addition, the minimum point of the DEP
Figure 5. a) Cyclic voltammetry tests voltage steps (top), current response (middle), and corresponding pressure response (bottom) upon cycling of
ASSBs using 7525-glass/C65 composite (red) and β-Li3PS4/C65 composite (blue) as working electrodes. Reproduced according to the terms of the
CC-BY license.[14g] Copyright 2021, The Authors, published by Wiley-VCH. b) The SEM images of the corresponding composites, 7525-glass/C65 (top)
and β-Li3PS4/C65 (bottom), after the CV tests. Reproduced with permission.[14g] Copyright 2021, The Authors, published by Wiley-VCH. c) CV profiles,
current response, and pressure response of Super C65 electrodes glassy SE (1.5Li2S–0.5P2S5–LiI) (left) and crystalline SE (Li6PS5Cl) (right). Reproduced
according to the terms of the CC-BY license.[94] Copyright 2022, The Authors, published by IOP Publishing Ltd. d) First-cycle charge voltage profiles and
corresponding pressure change (and DEP profiles) for poly-NCA electrodes employing LYC or LPSX. Reproduced with permission.[14f] Copyright 2021,
Wiley-VCH. e) First cycle charge-discharge voltage profiles and corresponding pressure changes for NCM electrodes using pristine and vulcanized BR
binder. Reproduced with permission.[14a] Copyright 2022, Elsevier.
Adv. Energy Mater. 2022, 2203153
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202203153 by Xiamen University, Wiley Online Library on [27/11/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advenergymat.de
© 2022 Wiley-VCH GmbH
2203153 (14 of 21)
www.advancedsciencenews.com
curve of P/LPSX (which can be regarded as an indicator of SOC
as discussed in Section4.2) shifts in the positive direction com-
pared with P/LYC, reflecting the delayed charging due to the
more severe decomposition reaction of LPSX (or Li+ deinterca-
lation). It is necessary to consider the electrochemo-mechanical
eect of SSEs as a design factor for the mechanical stability of
the electrode structure. A suitable SSE candidate should con-
sider its mechanical properties and the electrochemical stability
window.
In addition to considering the mechanical properties of SSEs
in the composite electrodes, the capability of the SSE sepa-
rator to resist the stress generated during the manufacturing
process and the volume changes of the electrode active mate-
rials during cycling are also closely related to their mechanical
properties.[95] An ideal SSE separator requires low porosity (to
achieve high ionic conductivity) as well as moderate elastic
modulus (including Young’s modulus (E), shear modulus (G),
bulk modulus (K), etc.) and fracture toughness, KIC. In gen-
eral, sulfide SSEs have favorable mechanical properties (lower
Pugh’s ratio, e.g., LPSCl with a G/B of 0.28, where G and B
are shear and bulk modulus, respectively) so that densification
of the electrolytes can be achieved at lower external pressures
(e.g., 360 MPa) and room temperature.[96] In comparison to
sulfide SSEs, another typical class of inorganic solid-state elec-
trolytes, oxide SSEs, due to distinctive dierences in mechan-
ical properties (e.g., Pugh’s ratio of 0.54 for LLZO), require a
much higher external pressure (up to hundreds of GPa) and
high temperature (up to 1000 °C) to achieve densification.[97] In
addition, the higher hardness (H) and elastic modulus (hard-
ness H= 9.2, 9.1, 7.1GPa, and Young’s modulus E= 200, 150,
115 GPa for LLTO, LLZO, and LATP, respectively) of oxide
SSEs than sulfide SSEs (detailed in Table 2) make them less
prone to elastic deformation.[11g] Therefore, it can be expected
that greater stress will be generated at the electrolyte/electrode
interface in pure oxide-based ASSBs, leading to more severe
electrochemo-mechanical issues.
Particularly, when Li metal is used as an anode in ASSBs,
the mechanical properties of the oxide and sulfide SSE will
exhibit dierent eects on the working conditions of the Li
metal anode. Monroe and Newman proposed a classical model
for elaborating the suppression of Li dendrite growth and SSE
mechanical parameters.[77] According to their model, dendrite
growth would be suppressed when the shear modulus of SSE is
greater than twice that of Li metal (4GPa). Therefore, the high
mechanical modulus of oxide SSE should suppress dendrite
growth, but the actual situation is complicated. However, the
growth of lithium dendrites has been reported in many oxide
SSEs (including LLZO, LLZTO, LATP, etc.), where the shear
modulus of LLZO (60GPa) is more than tenfold higher exceeds
that of Li metal, contrary to Monroe and Newman’s theory.[98]
This result suggests a more complex micro-short-circuit mecha-
nism in ASSBs. For sulfide SSEs, their Young’s modulus (e.g.,
E (LPSCl) = 22 GPa) is not much dierent from that of Li
metal (7.8 GPa), and the fracture toughness KIC is lower (e.g.,
KIC LPSCl = 0.23MPa m1/2), so the internal stress during bat-
tery cycling is likely to lead to a heterogeneous distribution of
lithium protrusions, which increases the local KIC of the SSE/
Li interface.[96] The increased local KIC will cause the cracks
to occur within the SSE layer, and then the lithium dendrites
will grow along the cracks, resulting in the short-circuit phe-
nomenon. For oxide SSEs, due to their relatively high Young’s
modulus (e.g., E (LLZO) = 175GPa) and KIC (e.g., KIC LLZO =
1.25 MPa m1/2), the internal stress of the battery may lead to
plastic flow of lithium rather than crack growth within the
SSE.[96] In this case, micro-short-circuit behavior in oxide SSEs
may be dominated by internal lithium nucleation (lithium den-
drites propagate through voids and grain boundaries with pref-
erential ion pathways) rather than the electrochemo-mechanical
damage/dendrite penetration behavior in sulfide SSEs.[99]
Scalable sheet-type electrodes are necessary for the prac-
tical application of sulfide-based ASSBs. Due to their key role
in maintaining microstructural integrity, binders are indis-
pensable in the preparation of flexible electrodes. Kwon etal.
reported a novel vulcanized butadiene rubber (BR) binder
with excellent mechanical properties, which can partially sup-
press the electromechanical eects during cycling, thereby
improving the electrochemical performance of batteries.[14a] The
authors monitored the pressure changes during the cycling of
NCM/LTO batteries with pristine and vulcanized BR binders.
As depicted in Figure5e, the overall pressure change trend is
consistent with the lattice volume change of Ni-rich layered
oxides upon the (de)lithiation process. However, with a lower
discharge depth, the cells using pristine BR binder showed
higher stress values than those using vulcanized BR binder.
The authors assumed that higher pressure change reflects
the formation of cracks and/or voids caused by electrochemo-
mechanical degradation. In addition, they quantified the crack
area fractions in the bulk and interfacial regions using cross-
section field-emission scanning electron microscopy backscat-
tered electron mode (FESEM-BSE) to examine the microstruc-
tural evolution of NCM electrodes with pristine and vulcanized
BR binders after the first charging and discharging process.
After the first charging, the crack area fractions in both bulk
and interfacial regions increased significantly in both batteries.
However, the area fraction of cracks in the interfacial region of
the battery using a pristine BR binder is higher than the vul-
canized BR binder. After the first discharging, the average area
fraction of cracks in the interfacial region of the battery using
a pristine BR binder is also slightly higher than that of the vul-
canized BR binder. In addition, the area fraction of cracks in
the bulk region showed similar values after the first charging
and discharging. These results suggest that the debonding of
the electrode/electrolyte interface has a more severe eect on
electrochemo-mechanical degradation than the contact loss
inside the electrode. This work indicates the possibility of com-
plementary analysis using FESEM-BSE and operando pressure
measurements to reveal the evolution of electro-mechanical
degradation.
4.4. Correlation of State of Charge (SOC) of the Batteries and
Changes of Stresses of the Batteries
Accurately estimating the state of charge (SOC) of the batteries
is crucial but also challenging. Many electrochemical techniques
based on measurable parameters (e.g., voltage, current, imped-
ances, and their dierentiates) have been developed to estimate
the SOC of batteries.[100] However, due to the fact that these tech-
Adv. Energy Mater. 2022, 2203153
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202203153 by Xiamen University, Wiley Online Library on [27/11/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advenergymat.de
© 2022 Wiley-VCH GmbH
2203153 (15 of 21)
www.advancedsciencenews.com
niques generally require measurements at low current densities
or at open circuit potential in order to exclude the eects of
polarizations, as well as complex computations, the SOC’s esti-
mation accuracy will decrease and become more complicated
as the battery ages. The correspondence between the battery
pressure and the volume change (related to the electrochemical
process) makes it possible to track SOC changes roughly but
also quickly. The sensitivity and simplicity of estimating bat-
tery SOC by pressure measurements have been demonstrated
in liquid batteries.[101] Due to the intrinsic rigidity, the pres-
sure changes in ASSBs will be more pronounced. Therefore,
pressure measurement can provide a new tool for estimating
the SOC of ASSBs. Jung et al. tracked the SOC of NCM/
graphite full cells by the operando dierential electrochemical
Figure 6. a–c) Charge discharge voltage profiles of each electrode, corresponding pressure change curves (ΔP), and DEP profiles (dP/dt) of Gr elec-
trodes for NCM/Gr cells with n/p ratios of a) 1.2, b) 1.4, and c) 1.7 at 2nd cycle, 0.1C, and 30 °C. Reproduced with permission.[14c] Copyright 2020,
Wiley-VCH. d–f) Schematic of stress compensation by using 55:45 wt% NCM-811:LCO cathode composite. Reproduced with permission.[14k] Copyright
2018, The Royal Society of Chemistry.
Adv. Energy Mater. 2022, 2203153
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202203153 by Xiamen University, Wiley Online Library on [27/11/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advenergymat.de
© 2022 Wiley-VCH GmbH
2203153 (16 of 21)
www.advancedsciencenews.com
pressiometry (DEP) method.[14c] Figure 6a–c shows the voltage
curves and pressure changes (ΔPGr) of NCM/graphite cells
with dierent n/p ratios. ΔPGr was estimated from the pressure
change in NCM/Gr battery minus that in NCM/LTO battery.
The results show that, as the n/p ratio increases, the minimum
DEP value (red diamonds in the figure) shifts in a positive direc-
tion during the charging process. The maximum point value
moves in the negative direction in the discharging process.
Because the total Gr electrode volume change is caused by Li+
intercalation (or deintercalation), this observation implies an
association between DEP profiles and specific SOCs. Compared
to traditional voltage-based estimation methods, pressure meas-
urements are more sensitive to SOC (especially for materials
with voltage plateaus during cycling, such as LFP), making it
more convenient to estimate SOC. In addition, since the pres-
sure can respond instantaneously to the lithium content in the
electrode, it can eectively detect the SOC dierence caused
by self-discharge compared to the traditional ampere-hour
counting method.
5. Summary and Outlook
Through the external pressure sensor or the built-in optical
FBG sensor, the internal stress change during the cycling
process of the solid-state battery can be obtained. This review
overviews the recent work on stress-measuring techniques and
their applications in sulfide-based ASSBs. It is expected that
stress(pressure)-measuring techniques in ASSBs, including
sulfide-based ASSBs, will have rapid development in the next
few years, and some future perspectives of the techniques are
outlined below.
1) The accuracy and resolution of pressure (or stress)-measur-
ing techniques need further improvement. For example, an
FSR pressure sensor with force resolution better than 0.01%
full scale is expected when studying active electrode mate-
rials with small volume changes (such as NCM materials).
In the pressure measuring experiment using the external
pressure sensor, the pressure change interferences caused by
the relaxation of mechanical components are not negligible.
Therefore, rational designs are required to suppress the me-
chanical noise and keep the pressure baseline as stable as
possible. In many pressure measurement experiments using
external pressure sensors, the actual measured pressure is
the addition of the total resultant force of the cathode and
anode stress changes, while the contributions of the stress
changes of these two electrodes are often inconsistent. There-
fore, it is necessary to decouple the cathode and anode stress
contributions separately. The stress decoupling methods for
electrodes composed of lithium-containing and lithium-free
active materials are dierent. For lithium-containing cathode
or anode active materials, such as NCM, LCO, Li2S, Li metal,
Li–In alloy, etc., the zero-strain LTO counter electrode can
be directly used to decouple the electrode stress. [14k,16,55] For
lithium-free active materials, such as S, graphite, Si, etc., it is
more dicult to decouple the stress changes of their corre-
sponding electrodes. Using the LTO electrode as a bridge is a
feasible method to deduct the stress contribution of the other
electrode from the stress change of the full cell (as reported
in an NCM-Gr battery[14c]). In addition, it has been proved
that decoupling of the pressure/volume change of a single
cathode electrode in liquid battery systems by using (elec-
tro)chemically prelithiated LTO as a counter electrode.[43,102]
However, its application in sulfide-based ASSBs remains to
be explored and verified. Comparing the dierential pressure
(dP/dV) and dierential capacity (dQ/dV) data is also needed
to correlate the electrochemical-stress relationship,[103] as
stress changes may not be apparent in the raw measurement
data.
2) Systematic local stress measurement of electrodes should be
emphasized, and it is still lacking in the literature. During
battery cycling, the uneven electrochemical reaction within
the electrode will lead to uneven local stress distribution and
result in serious electrochemo-mechanical degradation of the
batteries. Apparently, the local stress cannot be measured by
only a single external pressure sensor. Using an optical FBG
sensor is considered to be able to obtain local stress informa-
tion to a certain extent. In the follow-up experiments, more
sophisticated stress measuring methods, such as pressure
sensor arrays or the combination of multiple groups of op-
tical FBG sensors, may be able to obtain more detailed in-
formation on local stress distribution at the electrode/cell/
module level.
Due to their non-destructiveness and cost-friendliness,
external pressure sensor arrays have been applied to monitor
the local stress distribution of electrodes and the mechanical
distribution of battery modules in conventional liquid LIBs.[104]
As shown in Figure 7a, local stress mapping of anode lithium
deposition can be obtained using an external pressure sensor
array in a single cell.[105] The analysis of nonuniform stress dis-
tribution can provide some useful information for improving
local lithium deposition and detecting lithium dendrite growth.
In addition, in a battery module/pack composed of parallel-
series connected cells, the heterogeneity caused by the manu-
facturing process and the non-uniform working conditions will
lead to inconsistency between individual cells.[97] Inconsistency
between individual cells will lead to an uneven distribution of
internal stress in the module/pack, posing challenges to the
safety design and cyclic performance of the batteries. There-
fore, accurate monitoring of the stress distribution at the cell
level and mechanical inhomogeneity at the module/pack level
through the pressure sensor array is also essential for designing
the optimal mechanical structure and achieving superior elec-
trochemical performance (Figure7b).[106]
Implanted inside the cell, the optical FBG sensors can
provide local stress distribution within the electrodes and
between cell components. Local stress spatial resolution can
be achieved through quasi-distributed and fully distributed
FBG sensor designs. The design of the quasi-distributed FBG
sensor includes multiplexing of multiple FBG sensor groups
(Figure 7c) and integrating more than one grating with dif-
ferent Bragg wavelength (λB) in a single fiber by tuning the
Bragg grating period (Λ) (Figure 7d).[88b,104] Fully distrib-
uted FBG sensors based on Rayleigh, anti-Stokes Raman, or
Brillouin scattering can realize stress/temperature distribu-
Adv. Energy Mater. 2022, 2203153
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202203153 by Xiamen University, Wiley Online Library on [27/11/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advenergymat.de
© 2022 Wiley-VCH GmbH
2203153 (17 of 21)
www.advancedsciencenews.com
tion mapping in a single FBG sensor (Figure7e).[88b] Recently,
Huang et al. successfully obtained temperature distribution
mapping with a high spatial resolution of µm by a fully dis-
tributed FBG sensor based on Rayleigh scattering in a sodium-
ion battery.[88b] Although there have been no reports on stress
distribution mapping in batteries by fully distributed FBG sen-
sors, we believe that it is promising to achieve high-resolution
stress mapping by rationally designing experimental conditions
to eliminate the eects of temperature (T) and strain (ε) on λB
shift.[16]
For local stress information measurements of sulfide-based
ASSBs, the above schemes, which have been applied in liquid
batteries, can provide some guidance and implications. Trans-
planting external pressure sensor arrays from liquid to solid-
state batteries may not be so dicult. However, for built-in FBG
sensors, many new challenges, such as minimizing the impact
of FBG sensor implantation on the performance of solid-state
batteries and investigating chemical stability between FBG sen-
sors and sulfide SSEs, remain to be addressed.
3) The complexity of internal stress sources should be con-
sidered and analyzed more carefully. Taking layered oxide
materials as an example, NCM materials in liquid batteries
can cause gas evolution, which originates from electrolyte/
surface impurity decomposition and the release of lattice
oxygen at high voltage, mainly composed of CO2 and/or
O2. Interestingly, recent studies have shown that gas evolu-
tion due to side reactions can also be observed in sulfide-
based ASSBs.[107] Strauss et al. investigated the gas evolu-
tion of NCM622 and solid electrolyte (β-Li3PS4 or Li6PS5Cl)
by isotopic labeling and dierential electrochemical mass
spectrometry.[108] Results showed that CO2 and O2 could be
detected during the cycling process. Both species originate
from the electrochemical decomposition of Li2CO3 surface
contaminants, and O2 originates from oxygen evolution
from the NCM lattice at high SOC. This irreversible struc-
tural change of the active material is also one of the sources
of irreversible stress in the battery. Janek etal. first proposed
the concept of “stress balance” in a composite cathode for
sulfide-based all-solid-state battery. [14k] They proposed to
use NCM/LCO two active materials mixed in a 55:45 mass
ratio in the composite cathode. Due to the fact that the vol-
ume changes of NCM and LCO during the charging and
discharging process are opposite, the stress generated in the
composite cathode will cancel each other out (Figure6d–f).
However, considering that the internal stress changes in the
battery are derived from the stress coupling caused by the
cathode and anode, it is not enough to optimize the stress
balance of one electrode only. One possible approach is to
make sulfide-based ASSBs work in a designed stress-com-
pensating device. By externally keeping the pressure con-
stant, the electrochemo-mechanical issues will be alleviated.
In conclusion, although the stress (pressure)-measuring
technique is a powerful tool for detecting internal stress
changes directly, in order to correlate the relationship between
the electrochemical reactions and internal stress evolutions
in the batteries, a combination of several characterization
tools is still needed to understand and resolve these complex
electrochemo-mechanical issues in sulfide-based ASSBs. For
Figure 7. Schematic illustration of local stress distribution measurement in sulfide-based ASSBs. Reproduced with permission.[88b] Copyright 2021,
The Electrochemical Society. Reproduced with permission.[104] Copyright 2022, Elsevier. Reproduced with permission.[105] Copyright 2020, The Electro-
chemical Society. Reproduced with permission.[106] Copyright 2022, Elsevier.
Adv. Energy Mater. 2022, 2203153
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202203153 by Xiamen University, Wiley Online Library on [27/11/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advenergymat.de
© 2022 Wiley-VCH GmbH
2203153 (18 of 21)
www.advancedsciencenews.com
example, combining pressure-probe with in-situ X-ray dirac-
tion (XRD), X-ray absorption spectroscopy (XAS), or solid-
state NMR can obtain information on stress changes and local
structural/valence changes of active materials, and imaging
methods such as SEM or CT can more intuitively observe the
relationship between stress change and battery level structure
change. In addition, computational mechanical models of the
solid-state batteries constructed by phase-field methods will
be helpful in analyzing experimental data on the stresses and
predicting more complex electrochemo-mechanical pheno-
mena during the charging/discharging processes of the
batteries.
Acknowledgements
This work was financially supported by National Natural Science
Foundation of China (Grant No. 21935009) and the National Key R&D
Program of China (Grant No. 2021YFB2401800).
Conflict of Interest
The authors declare no conflict of interest.
Keywords
electrochemo-mechanical eects, electroche-mechanical stresses, stress
measurement, sulfide-based ASSBs
Received: September 16, 2022
Revised: November 2, 2022
Published online:
[1] a) J. M. Tarascon, M. Armand, Nature 2001, 414, 359;
b) A. Manthiram, Nat. Commun. 2020, 11, 1550; c) S. Chu,
A.Majumdar, Nature 2012, 488, 294; d) M.Armand, J. M.Tarascon,
Nature 2008, 451, 652.
[2] a) J. B. Goodenough, Y. Kim, Chem. Mater. 2010, 22, 587;
b) Q. Wang, P.Ping, X.Zhao, G. Chu, J. Sun, C. Chen, J. Power
Sources 2012, 208, 210.
[3] a) X. Zhang, Y. Yang, Z. Zhou, Chem. Soc. Rev. 2020, 49, 3040;
b) W. Xu, J. Wang, F. Ding, X. Chen, E. Nasybulin, Y. Zhang,
J.-G.Zhang, Energy Environ. Sci. 2014, 7, 513.
[4] a) X.-B.Cheng, R. Zhang, C.-Z.Zhao, Q.Zhang, Chem. Rev. 2017,
117, 10403; b) Y. Zhang, T.-T. Zuo, J. Popovic, K. Lim, Y.-X. Yin,
J.Maier, Y.-G.Guo, Mater. Today 2020, 33, 56.
[5] a) A.Manthiram, X.Yu, S.Wang, Nat. Rev. Mater. 2017, 2, 16103;
b) Y.-S. Hu, Nat. Energy 2016, 1, 16042; c) L. Fan, S. Wei, S. Li,
Q.Li, Y.Lu, Adv. Energy Mater. 2018, 8, 1702657; d) G. Cui, Matter
2020, 2, 805.
[6] a) M. Dixit, A. Parejiya, R. Essehli, N.Muralidharan, S. U.Haq,
R. Amin, I. Belharouak, Cell. Rep. Phys. Sci. 2022, 3, 100756;
b) T.Liu, Y.Yuan, X.Tao, Z.Lin, J.Lu, Adv. Sci. 2020, 7, 2001207.
[7] a) Y.Liang, H. Liu, G.Wang, C.Wang, Y. Ni, C. W.Nan, L. Z. Fan,
InfoMat 2022, 4, e12292; b) J.Wu, S.Liu, F.Han, X.Yao, C.Wang,
Adv. Mater. 2021, 33, 2000751; c) Q. Zhang, D. Cao, Y. Ma,
A.Natan, P.Aurora, H.Zhu, Adv. Mater. 2019, 31, 1901131.
[8] a) A.Mukhopadhyay, B. W.Sheldon, Prog. Mater. Sci. 2014, 63, 58;
b) Y.Li, X.Cheng, Y.Zhang, K.Zhao, Mater. Today Nano 2019, 7,
100040; c) L.Xu, W.Liu, Y. Hu, L.Luo, Chem. Commun. 2020, 56,
13301.
[9] a) R.Xu, K. Zhao, J. Electrochem. Energy Convers. Storage 2016, 13,
030803; b) Y. Zhao, P.Stein, Y. Bai, M. Al-Siraj, Y.Yang, B.-X.Xu,
J. Power Sources 2019, 413, 259.
[10] a) M.-C. Pang, K. Yang, R. Brugge, T. Zhang, X. Liu, F. Pan,
S. Yang, A. Aguadero, B. Wu, M. Marinescu, Mater. Today 2021,
49, 145; b) J. A.Lewis, J. Tippens, F. J. Q.Cortes, M. T.McDowell,
Trends Chem. 2019, 1, 845; c) T. Shi, Y.-Q. Zhang, Q.Tu, Y.Wang,
M. C.Scott, G.Ceder, J. Mater. Chem. A 2020, 8, 17399; d) J.Tian,
Z. Chen, Y. Zhao, ACS Omega 2022, 7, 6455; e) J.Liu, H.Yuan,
H.Liu, C. Z.Zhao, Y.Lu, X. B.Cheng, J. Q.Huang, Q.Zhang, Adv.
Energy Mater. 2022, 12, 2100748; f ) Y.Ren, K. B. Hatzell, J. Mater.
Chem. A 2021, 9, 13804.
[11] a) L. R.Mangani, C. Villevieille, J. Mater. Chem. A 2020, 8, 10150;
b) R. Koerver, I. Aygün, T. Leichtweiß, C. Dietrich, W. Zhang,
J. O. Binder, P.Hartmann, W. G.Zeier, J. R.Janek, Chem. Mater.
2017, 29, 5574; c) S. Ohno, R.Koerver, G.Dewald, C. Rosenbach,
P. Titscher, D. Steckermeier, A. Kwade, J. r. Janek, W. G. Zeier,
Chem. Mater. 2019, 31, 2930; d) T. Krauskopf, F. H. Richter,
W. G. Zeier, J. r. Janek, Chem. Rev. 2020, 120, 7745; e) X. Yao,
N. Huang, F. Han, Q.Zhang, H.Wan, J. P. Mwizerwa, C. Wang,
X.Xu, Adv. Energy Mater. 2017, 7, 1602923; f ) G.Bucci, T.Swamy,
Y.-M. Chiang, W. C. Carter, J. Mater. Chem. A 2017, 5, 19422;
g) F.Zhang, Q.-A.Huang, Z.Tang, A.Li, Q.Shao, L.Zhang, X. Li,
J.Zhang, Nano Energy 2020, 70, 104545.
[12] a) N.Sun, Q.Liu, Y.Cao, S.Lou, M.Ge, X.Xiao, W. K.Lee, Y.Gao,
G.Yin, J.Wang, Angew. Chem. 2019, 131, 18820; b) X.Liu, B.Zheng,
J.Zhao, W.Zhao, Z.Liang, Y.Su, C.Xie, K.Zhou, Y.Xiang, J.Zhu,
Adv. Energy Mater. 2021, 11, 2003583.
[13] M. R. Busche, D. A.Weber, Y. Schneider, C. Dietrich, S. Wenzel,
T. Leichtweiss, D. Schröder, W. Zhang, H. Weigand, D. Walter,
S. J. Sedlmaier, D. Houtarde, L. F.Nazar, J. Janek, Chem. Mater.
2016, 28, 6152.
[14] a) T. Y.Kwon, K. T. Kim, D. Y. Oh, Y. B. Song, S.Jun, Y. S.Jung,
Energy Storage Mater. 2022, 49, 219; b) S. Y.Han, C.Lee, J. A.Lewis,
D. Yeh, Y. Liu, H.-W. Lee, M. T. McDowell, Joule 2021, 5, 2450;
c) S.Jun, Y. J.Nam, H.Kwak, K. T.Kim, D. Y.Oh, Y. S. Jung, Adv.
Funct. Mater. 2020, 30, 2002535; d) S. W.Park, H. J.Choi, Y.Yoo,
H. D. Lim, J. W. Park, Y. J. Lee, Y. C. Ha, S. M. Lee, B. G. Kim,
Adv. Funct. Mater. 2022, 32, 2108203; e) S. Cangaz, F. Hippauf,
F. S. Reuter, S. Doerfler, T.Abendroth, H. Althues, S.Kaskel, Adv.
Energy Mater. 2020, 10, 2001320; f ) Y. Han, S. H.Jung, H.Kwak,
S.Jun, H. H.Kwak, J. H. Lee, S. T.Hong, Y. S.Jung, Adv. Energy
Mater. 2021, 11, 2100126; g) S.Wang, W.Zhang, X.Chen, D.Das,
R.Ruess, A.Gautam, F.Walther, S.Ohno, R. Koerver, Q.Zhang,
Adv. Energy Mater. 2021, 11, 2100654; h) X.Li, J. Liang, J. T. Kim,
J.Fu, H.Duan, N.Chen, R.Li, S.Zhao, J.Wang, H.Huang, X.Sun,
Adv. Mater. 2022, 34, 2200856; i) S. Wang, M.Tang, Q. Zhang,
B. Li, S. Ohno, F. Walther, R. Pan, X. Xu, C. Xin, W. Zhang,
Adv. Energy Mater. 2021, 11, 2101370; j) W. Zhang, D. Schröder,
T. Arlt, I. Manke, R. Koerver, R. Pinedo, D. A. Weber, J. Sann,
W. G.Zeier, J.Janek, J. Mater. Chem. A 2017, 5, 9929; k) R.Koerver,
W. Zhang, L.de Biasi, S.Schweidler, A. O.Kondrakov, S.Kolling,
T. Brezesinski, P. Hartmann, W. G.Zeier, J.Janek, Energy Environ.
Sci. 2018, 11, 2142; l) L. Peng, C. Yu, Z. Zhang, R. Xu, M. Sun,
L.Zhang, S. Cheng, J.Xie, Energy Environ. Mater. 2021, 5, https://
doi.org/10.1002/eem2.12308; m) C. Lee, S. Y. Han, J. A. Lewis,
P. P.Shetty, D. Yeh, Y.Liu, E. Klein, H.-W.Lee, M. T. McDowell,
ACS Energy Lett. 2021, 6, 3261.
[15] a) G. Yang, C. Leitão, Y.Li, J.Pinto, X. Jiang, Measurement 2013,
46, 3166; b) L. W. Sommer, P. Kiesel, A. Ganguli, A. Lochbaum,
B.Saha, J.Schwartz, C.-J.Bae, M.Alamgir, A.Raghavan, J. Power
Sources 2015, 296, 46; c) M.Nascimento, S. Novais, M. S. Ding,
M. S. Ferreira, S.Koch, S. Passerini, J. L. Pinto, J. Power Sources
Adv. Energy Mater. 2022, 2203153
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202203153 by Xiamen University, Wiley Online Library on [27/11/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advenergymat.de
© 2022 Wiley-VCH GmbH
2203153 (19 of 21)
www.advancedsciencenews.com
2019, 410, 1; d) J. Peng, X. Zhou, S. Jia, Y. Jin, S. Xu, J. Chen,
J. Power Sources 2019, 433, 226692; e) P.Desai, J.Huang, H.Hijazi,
L.Zhang, S. Mariyappan, J. M.Tarascon, Adv. Energy Mater. 2021,
11, 2101490; f) M. Nascimento, M. S.Ferreira, J. L.Pinto, Batteries
2018, 4, 23; g) M.Nascimento, T.Paixão, M. S.Ferreira, J. L.Pinto,
Batteries 2018, 4, 67.
[16] L.Albero Blanquer, F.Marchini, J. R. Seitz, N.Daher, F.Bétermier,
J.Huang, C.Gervillié, J.-M.Tarascon, Nat. Commun. 2022, 13, 1153.
[17] Y.Bae Song, H. Kwak, W. Cho, K. S.Kim, Y.Seok Jung, K.-H.Park,
Curr. Opin. Solid State Mater. Sci. 2022, 26, 100977.
[18] a) P. Barai, K.Higa, A. T.Ngo, L. A.Curtiss, V.Srinivasan, J. Elec-
trochem. Soc. 2019, 166, A1752; b) R. M.McMeeking, M. Ganser,
M. Klinsmann, F. E. Hildebrand, J. Electrochem. Soc. 2019, 166,
A984; c) H.-K.Tian, A.Chakraborty, A. A.Talin, P.Eisenlohr, Y.Qi,
J. Electrochem. Soc. 2020, 167, 090541; d) Y. He, C. Lu, S. Liu,
W.Zheng, J.Luo, Adv. Energy Mater. 2019, 9, 1901810.
[19] A.Mistry, P. P.Mukherjee, J. Electrochem. Soc. 2020, 167, 082510.
[20] J.Xu, F.Lin, M. M.Doe, W.Tong, J. Mater. Chem. A 2017, 5, 874.
[21] P.Rozier, J. M.Tarascon, J. Electrochem. Soc. 2015, 162, A2490.
[22] G. G. Amatucci, J. M. Tarascon, L. C. Klein, J. Electrochem. Soc.
1996, 143, 1114.
[23] a) J.Cho, Y. J. Kim, B.Park, J. Electrochem. Soc. 2001, 148, A1110;
b) J. N.Reimers, J. R.Dahn, J. Electrochem. Soc. 1992, 139, 2091;
c) O. Dolotko, A. Senyshyn, M. J. Mühlbauer, K. Nikolowski,
H.Ehrenberg, J. Power Sources 2014, 255, 197.
[24] M.Bianchini, M.Roca-Ayats, P.Hartmann, T.Brezesinski, J.Janek,
Angew. Chem., Int. Ed. 2019, 58, 10434.
[25] L. de Biasi, A. Schiele, M.Roca-Ayats, G.Garcia, T. Brezesinski,
P.Hartmann, J.Janek, ChemSusChem 2019, 12, 2240.
[26] P. H. Chien, X.Wu, B.Song, Z.Yang, C. K.Waters, M. S.Everett,
F.Lin, Z.Du, J.Liu, Batteries Supercaps 2021, 4, 1701.
[27] a) R. Robert, C.Villevieille, P. Novák, J. Mater. Chem. A 2014, 2,
8589; b) P.He, H.Yu, H.Zhou, J. Mater. Chem. 2012, 22, 3680.
[28] F.Strauss, L.deBiasi, A. Y.Kim, J.Hertle, S.Schweidler, J. r.Janek,
P.Hartmann, T.Brezesinski, ACS Mater. Lett. 2019, 2, 84.
[29] X.Wang, H.Zhou, Z. Chen, X.Meng, Energy Storage Mater. 2022,
49, 181.
[30] H.-H.Ryu, K.-J.Park, C. S. Yoon, Y.-K.Sun, Chem. Mater. 2018, 30,
1155.
[31] L.Liu, M. Li, L.Chu, B.Jiang, R.Lin, X.Zhu, G. Cao, Prog. Mater.
Sci. 2020, 111, 100655.
[32] G. W. Nam, N.-Y. Park, K.-J. Park, J. Yang, J. Liu, C. S. Yoon,
Y.-K.Sun, ACS Energy Lett. 2019, 4, 2995.
[33] A.-H.Marincaş, F.Goga, S.-A.Dorneanu, P.Ilea, J. Solid State Elec-
trochem. 2020, 24, 473.
[34] a) K.Kanamura, H.Naito, T.Yao, Z.-I.Takehara, J. Mater. Chem. 1996,
6, 33; b) T.Ohzuku, M.Kitagawa, T.Hirai, J. Electrochem. Soc. 1990,
137, 769; c) Y. Qi, L. G.Hector, C. James, K. J.Kim, J. Electrochem.
Soc. 2014, 161, F3010; d) X.Sun, X. Q.Yang, M.Balasubramanian,
J. McBreen, Y. Xia, T. Sakai, J. Electrochem. Soc. 2002, 149, A842;
e) Y.Zhang, H.Xie, H.Jin, Q.Zhang, Y. Li, X.Li, K. Li, C.Bao, IOP
Conf. Ser.: Earth Environ. Sci. 2020, 603, 012051.
[35] S. Lee, Y. Oshima, E. Hosono, H. Zhou, K. Kim, H. M. Chang,
R.Kanno, K.Takayanagi, J. Phys. Chem. C 2013, 117, 24236.
[36] Y. Koyama, T. E. Chin, U. Rhyner, R. K. Holman, S. R. Hall,
Y. M.Chiang, Adv. Funct. Mater. 2006, 16, 492.
[37] A. K. Padhi, K. S.Nanjundaswamy, J. B. Goodenough, J. Electro-
chem. Soc. 1997, 144, 1188.
[38] a) W.-J. Zhang, J. Power Sources 2011, 196, 2962; b) I. A.Bobrikov,
A. M. Balagurov, C.-W. Hu, C.-H. Lee, T.-Y. Chen, S. Deleg,
D. A.Balagurov, J. Power Sources 2014, 258, 356; c) Q.Liu, H.He,
Z.-F. Li, Y. Liu, Y.Ren, W.Lu, J.Lu, E. A. Stach, J.Xie, ACS Appl.
Mater. Interfaces 2014, 6, 3282; d) L.Gu, C.Zhu, H.Li, Y. Yu, C.Li,
S. Tsukimoto, J.Maier, Y. Ikuhara, J. Am. Chem. Soc. 2011, 133,
4661; e) Y.Orikasa, T.Maeda, Y.Koyama, H.Murayama, K.Fukuda,
H.Tanida, H.Arai, E. Matsubara, Y. Uchimoto, Z.Ogumi, J. Am.
Chem. Soc. 2013, 135, 5497.
[39] a) F. Wu, G. Yushin, Energy Environ. Sci. 2017, 10, 435;
b) M. Chen, R. P. Rao, S. Adams, Solid State Ionics 2014, 268,
300; c) S. M. Hosseini, A. Varzi, S. Ito, Y. Aihara, S. Passerini,
Energy Storage Mater. 2020, 27, 61; d) J. P.Mwizerwa, Q.Zhang,
F. Han, H. Wan, L. Cai, C.Wang, X. Yao, ACS Appl. Mater. Inter-
faces 2020, 12, 18519; e) A. L. Santhosha, N. Nazer, R. Koerver,
S. Randau, F. H.Richter, D. A. Weber, J. Kulisch, T. Adermann,
J. Janek, P. Adelhelm, Adv. Energy Mater. 2020, 10, 2002394;
f) Q. Zhang, J. P. Mwizerwa, H. Wan, L. Cai, X. Xu, X. Yao, J.
Mater. Chem. A 2017, 5, 23919; g) U.Ulissi, S.Ito, S. M.Hosseini,
A. Varzi, Y. Aihara, S. Passerini, Adv. Energy Mater. 2018, 8,
1801462.
[40] S.Kim, J.Choi, S. M.Bak, L.Sang, Q.Li, A.Patra, P. V.Braun, Adv.
Funct. Mater. 2019, 29, 1901719.
[41] W. P.Wang, J.Zhang, J.Chou, Y. X. Yin, Y.You, S.Xin, Y. G.Guo,
Adv. Energy Mater. 2021, 11, 2000791.
[42] a) A.Mukhopadhyay, A.Tokranov, K.Sena, X.Xiao, B. W.Sheldon,
Carbon 2011, 49, 2742; b) S.Flandrois, B.Simon, Carbon 1999, 37,
165; c) A. Mukhopadhyay, A. Tokranov, X. Xiao, B. W. Sheldon,
Electrochim. Acta 2012, 66, 28.
[43] S.Schweidler, L.de Biasi, A.Schiele, P.Hartmann, T.Brezesinski,
J. r.Janek, J. Phys. Chem. C 2018, 122, 8829.
[44] a) B.Zhao, R.Ran, M. Liu, Z.Shao, Mater. Sci. Eng., R 2015, 98,
1; b) Q.Zhang, C. Zhang, B.Li, D. Jiang, S.Kang, X. Li, Y.Wang,
Electrochim. Acta 2013, 107, 139.
[45] J. Haetge, P. Hartmann, K. Brezesinski, J. Janek, T. Brezesinski,
Chem. Mater. 2011, 23, 4384.
[46] T. Ohzuku, A. Ueda, N. Yamamoto, J. Electrochem. Soc. 1995, 142,
1431.
[47] S. H. Jung, U. H.Kim, J. H.Kim, S. Jun, C. S. Yoon, Y. S. Jung,
Y. K.Sun, Adv. Energy Mater. 2020, 10, 1903360.
[48] M. T.McDowell, S. W. Lee, W. D.Nix, Y.Cui, Adv. Mater. 2013, 25,
4966.
[49] a) U.Kasavajjula, C. Wang, A. J. Appleby, J. Power Sources 2007,
163, 1003; b) J. Li, J. R. Dahn, J. Electrochem. Soc. 2007, 154,
A156.
[50] M. T. McDowell, S. W. Lee, J. T. Harris, B. A. Korgel, C. Wang,
W. D.Nix, Y.Cui, Nano Lett. 2013, 13, 758.
[51] T.Yoon, C. C.Nguyen, D. M. Seo, B. L.Lucht, J. Electrochem. Soc.
2015, 162, A2325.
[52] a) M.Nagao, A.Hayashi, M. Tatsumisago, Electrochemistry 2012,
80, 734; b) A. L. Santhosha, L. Medenbach, J. R. Buchheim,
P.Adelhelm, Batteries Supercaps 2019, 2, 524.
[53] a) D.Cao, Y.Zhang, A. M.Nolan, X.Sun, C.Liu, J.Sheng, Y.Mo,
Y. Wang, H.Zhu, Nano Lett. 2019, 20, 1483; b) S.Luo, Z. Wang,
X. Li, X.Liu, H. Wang, W. Ma, L.Zhang, L.Zhu, X. Zhang, Nat.
Commun. 2021, 12, 1.
[54] J. Liu, Z. Bao, Y. Cui, E. J. Dufek, J. B. Goodenough,
P. Khalifah, Q. Li, B. Y. Liaw, P. Liu, A. Manthiram,
Y. S. Meng, V. R.Subramanian, M. F. Toney, V. V.Viswanathan,
M. S.Whittingham, J.Xiao, W.Xu, J.Yang, X.-Q.Yang, J.-G.Zhang,
Nat. Energy 2019, 4, 180.
[55] P. Liang, G.Shao, H. Wang, C.-A.Wang, ACS Appl. Energy Mater.
2021, 4, 3993.
[56] K. H.Park, Q. Bai, D. H.Kim, D. Y.Oh, Y.Zhu, Y.Mo, Y. S. Jung,
Adv. Energy Mater. 2018, 8, 1800035.
[57] X.Yang, J.Luo, X.Sun, Chem. Soc. Rev. 2020, 49, 2140.
[58] C. Yu, J. Hageman, S. Ganapathy, L. van Eijck, L. Zhang,
K. R. Adair, X. Sun, M. Wagemaker, J. Mater. Chem. A 2019, 7,
10412.
[59] W. D. Richards, L. J. Miara, Y. Wang, J. C. Kim, G. Ceder, Chem.
Mater. 2016, 28, 266.
[60] J.Choi, A.Manthiram, J. Electrochem. Soc. 2005, 152, A1714.
Adv. Energy Mater. 2022, 2203153
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202203153 by Xiamen University, Wiley Online Library on [27/11/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advenergymat.de
© 2022 Wiley-VCH GmbH
2203153 (20 of 21)
www.advancedsciencenews.com
[61] L. de Biasi, A. O. Kondrakov, H. Geßwein, T. Brezesinski,
P.Hartmann, J. r.Janek, J. Phys. Chem. C 2017, 121, 26163.
[62] W.-S. Yoon, K. Y. Chung, J. McBreen, X.-Q. Yang, Electrochem.
Commun. 2006, 8, 1257.
[63] Y.Qi, H.Guo, L. G.Hector, A.Timmons, J. Electrochem. Soc. 2010,
157, A558.
[64] L. Y.Beaulieu, K. W.Eberman, R. L.Turner, L. J.Krause, J. R.Dahn,
Electrochem. Solid-State Lett. 2001, 4, A137.
[65] W. Zhou, S. Upreti, M. S. Whittingham, Electrochem. Commun.
2011, 13, 158.
[66] A.Schiele, T.Hatsukade, B. B.Berkes, P.Hartmann, T.Brezesinski,
J. r.Janek, Anal. Chem. 2017, 89, 8122.
[67] Y. B.Song, H.Kwak, W.Cho, K. S.Kim, Y. S.Jung, K.-H.Park, Curr.
Opin. Solid State Mater. Sci. 2022, 26, 100977.
[68] B.Noble, S. J.Harris, K.Dinsdale, J. Mater. Sci. 1982, 17, 461.
[69] E. J. Cheng, N. J. Taylor, J. Wolfenstine, J. Sakamoto, J. Asian
Ceram. Soc. 2017, 5, 113.
[70] E. J. Cheng, K. Hong, N. J. Taylor, H. Choe, J. Wolfenstine,
J.Sakamoto, J. Eur. Ceram. Soc. 2017, 37, 3213.
[71] T.Maxisch, G.Ceder, Phys. Rev. B 2006, 73, 174112.
[72] G. A.Saunders, Y. K.Yog¯urtçu, J. E.Macdonald, G. S.Pawley, Proc.
R. Soc. A 1986, 407, 325.
[73] H. Khachai, R. Khenata, A. Bouhemadou, A. H. Reshak,
A.Haddou, M.Rabah, B.Soudini, Solid State Commun. 2008, 147,
178.
[74] T.Lorenz, D.Teich, J.-O.Joswig, G.Seifert, J. Phys. Chem. C 2012,
116, 11714.
[75] Z.Deng, Z.Wang, I.-H.Chu, J.Luo, S. P.Ong, J. Electrochem. Soc.
2015, 163, A67.
[76] A. Kato, M.Yamamoto, A. Sakuda, A.Hayashi, M.Tatsumisago,
ACS Appl. Energy Mater. 2018, 1, 1002.
[77] C.Monroe, J.Newman, J. Electrochem. Soc. 2005, 152, A396.
[78] H. Ledbetter, N. Sizova, S. Kim, H. Kobayashi, S. Sgobba,
L.Parrini, J. Phys. IV 1996, 6, C8.
[79] V. B.Shenoy, P.Johari, Y.Qi, J. Power Sources 2010, 195, 6825.
[80] M. E.Stournara, P. R.Guduru, V. B.Shenoy, J. Power Sources 2012,
208, 165.
[81] T.-F.Yi, Y.Xie, Y.-R. Zhu, J.Shu, A.-N.Zhou, H.-B. Qiao, J. Power
Sources 2012, 198, 318.
[82] H. Kim, G.Kim, T.Kim, S.Lee, D. Kang, M. S.Hwang, Y.Chae,
S.Kang, H.Lee, H. G.Park, Small 2018, 14, 1703432.
[83] Z. Liu, X. Gu, C. Wu, H. Ren, Z. Zhou, S. Tang, Constr. Build.
Mater. 2022, 321, 126085.
[84] V. Kumar, A. Kumar, S. S. Han, S.-S.Park, Nano Mater. Sci. 2021,
3, 233.
[85] a) G. Canavese, S. Stassi, C. Fallauto, S. Corbellini, V. Cauda,
V.Camarchia, M. Pirola, C. F.Pirri, Sens. Actuators, A 2014, 208, 1;
b) N.Luo, W.Dai, C.Li, Z.Zhou, L.Lu, C. C. Y.Poon, S. C.Chen,
Y.Zhang, N.Zhao, Adv. Funct. Mater. 2016, 26, 1178.
[86] K. M. B.Jansen, Exp. Mech. 1997, 37, 245.
[87] A. H. A.Razak, A. Zayegh, R. K.Begg, Y. Wahab, Sensors 2012, 12,
9884.
[88] a) Z.Miao, Y.Li, X.Xiao, Q.Sun, B.He, X.Chen, Y.Liao, Y.Zhang,
L.Yuan, Z.Yan, Z.Li, Y.Huang, Energy Environ. Sci. 2022, 15, 2029;
b) J.Huang, L. A. Blanquer, C.Gervillié, J.-M.Tarascon, J. Electro-
chem. Soc. 2021, 168, 060520.
[89] J.Huang, L. Albero Blanquer, J. Bonefacino, E. R.Logan, D.Alves
Dalla Corte, C.Delacourt, B. M.Gallant, S. T.Boles, J. R.Dahn,
H.-Y.Tam, Nat. Energy 2020, 5, 674.
[90] J. A.Lewis, C.Lee, Y.Liu, S. Y.Han, D.Prakash, E. J.Klein, H.-W.Lee,
M. T.McDowell, ACS Appl. Mater. Interfaces 2022, 14, 4051.
[91] B. A.Budiman, A. Saputro, S.Rahardian, M.Aziz, P.Sambegoro,
I. P.Nurprasetio, J. Energy Storage 2022, 52, 104810.
[92] H.Wang, J. Zhu, Y.Su, Z.Gong, Y. Yang, Sci. China Chem. 2021,
64, 879.
[93] H. Lim, S.Jun, Y. B. Song, H. Bae, J. H. Kim, Y. S. Jung, Energy
Storage Mater. 2022, 50, 543.
[94] J. H. Teo, F.Strauss, F. Walther, Y. Ma, S. Payandeh, T. Scherer,
M. Bianchini, J. Janek, T. Brezesinski, Mater. Futures 2022, 1,
015102.
[95] a) J.Schnell, T.Günther, T. Knoche, C.Vieider, L. Köhler, A. Just,
M.Keller, S.Passerini, G.Reinhart, J. Power Sources 2018, 382, 160;
b) J.Liang, J.Luo, Q. Sun, X.Yang, R.Li, X.Sun, Energy Storage
Mater. 2019, 21, 308.
[96] D. K. Singh, A. Henss, B. Mogwitz, A. Gautam, J. Horn,
T.Krauskopf, S.Burkhardt, J.Sann, F. H.Richter, J.Janek, Cell Rep.
Phys. Sci. 2022, 3, 101043.
[97] R.Li, W.Li, A. Singh, D.Ren, Z.Hou, M.Ouyang, Energy Storage
Mater. 2022, 52, 395.
[98] a) J. Zhu, J.Zhao, Y. Xiang, M. Lin, H. Wang, B.Zheng, H. He,
Q. Wu, J. Y. Huang, Y. Yang, Chem. Mater. 2020, 32, 4998;
b) L.Porz, T.Swamy, B. W.Sheldon, D.Rettenwander, T.Frömling,
H. L.Thaman, S. Berendts, R.Uecker, W. C. Carter, Y. M. Chiang,
Adv. Energy Mater. 2017, 7, 1701003; c) H.Wang, H.Gao, X.Chen,
J. Zhu, W.Li, Z. Gong, Y. Li, M. S. Wang, Y. Yang, Adv. Energy
Mater. 2021, 11, 2102148.
[99] H. Shen, K. Chen, J. Kou, Z. Jia, N. Tamura, W. Hua, W. Tang,
H.Ehrenberg, M.Doe, Mater. Today 2022, 57, 190.
[100] a) A. Barré, B. Deguilhem, S. Grolleau, M. Gérard, F. Suard,
D.Riu, J. Power Sources 2013, 241, 680; b) J.Zhang, J.Lee, J. Power
Sources 2011, 196, 6007; c) J.Remmlinger, M.Buchholz, M.Meiler,
P. Bernreuter, K. Dietmayer, J. Power Sources 2011, 196, 5357;
d) K.Maher, R.Yazami, J. Power Sources 2014, 247, 527; e) X.Han,
M.Ouyang, L.Lu, J.Li, Y.Zheng, Z.Li, J. Power Sources 2014, 251,
38; f) J. Remmlinger, M.Buchholz, T.Soczka-Guth, K.Dietmayer,
J. Power Sources 2013, 239, 689.
[101] a) L. De Sutter, G. Berckmans, M. Marinaro, M. Wohlfahrt-
Mehrens, M. Berecibar, J. Van Mierlo, J. Power Sources 2020,
451, 227774; b) H. Dai, C. Yu, X.Wei, Z. Sun, Energy 2017, 129,
16; c) J. Cannarella, C. B. Arnold, J. Power Sources 2014, 269,
7; d) M. A. Figueroa-Santos, J. B. Siegel, A. G. Stefanopoulou,
Energies 2020, 13, 2653; e) P. Mohtat, S. Lee, J. B. Siegel,
A. G.Stefanopoulou, J. Electrochem. Soc. 2021, 168, 100520.
[102] a) M. Göttlinger, P. Daubinger, W. Stracke, S. Hartmann,
G. A. Gin, Electrochim. Acta 2022, 419, 140354;
b) P. R. Chinnam, L.Xu, L. Cai, N. L.Cordes, S.Kim, C. M. Efaw,
D. J.Murray, E. J.Dufek, H.Xu, B.Li, Adv. Energy Mater. 2022, 12,
2103048.
[103] S. Kim, A. Raj, B. Li, E. J. Dufek, C. C.Dickerson, H.-Y. Huang,
B.Liaw, G. M.Pawar, J. Power Sources 2020, 463, 228180.
[104] X.Lu, J.-M.Tarascon, J.Huang, eTransportation 2022, 14, 100207.
[105] L.Hovestadt, K.Wildermann, A.Sahhary, R.Hanke-Rauschenbach,
J. Electrochem. Soc. 2020, 167, 110540.
[106] X.Zhong, L.Yang, N.Li, Z.Chu, J.Chen, S.Zhu, W.-L.Song, S.Ai,
H.-S.Chen, J. Power Sources 2022, 548, 232053.
[107] T. Bartsch, F. Strauss, T. Hatsukade, A. Schiele, A. Y. Kim,
P.Hartmann, J. r.Janek, T.Brezesinski, ACS Energy Lett. 2018, 3, 2539.
[108] F. Strauss, J. H. Teo, A. Schiele, T. Bartsch, T. Hatsukade,
P.Hartmann, J. r.Janek, T.Brezesinski, ACS Appl. Mater. Interfaces
2020, 12, 20462.
[109] A. Fortier, M.Tsao, N. D.Williard, Y. Xing, M. G. Pecht, Energies
2017, 10, 838.
Adv. Energy Mater. 2022, 2203153
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202203153 by Xiamen University, Wiley Online Library on [27/11/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advenergymat.de
© 2022 Wiley-VCH GmbH
2203153 (21 of 21)
www.advancedsciencenews.com
Jiabao Gu is a Ph.D. candidate at Xiamen University under the supervision of Prof. Yong Yang. His
current research topic focuses on the electrochemo-mechanical issues in all-solid-state batteries.
Ziteng Liang received his Bachelor’s degree in chemistry from Xiamen University in 2018. He
is now pursuing his Ph.D. degree in Collaborative Innovation Center of Chemistry for Energy
Materials in Xiamen University, under the supervision of Prof. Yong Yang. His research is dedi-
cated to investigating the failure mechanisms of sulfide-based solid-state batteries and presenting
corresponding solutions.
Yong Yang is now working as a distinguished professor at Xiamen University. He also serves
as Editor for J. Power Sources and The First Vice-President of Board Member of International
Battery Materials Association (IBA) and International Meeting of Lithium Battery (IMLB). His
main research interests are electrode/electrolyte materials for Li/Na-ion batteries, Solid State
Battery, and Battery Degradation Analysis of Li-ion Battery etc. He has published 400+ papers in
referred journals and got 20000+ citations (Goggle Scholar, H = 75). He is the recipient of 2020
Technology Award and Battery Division, ECS and Technology Award presented by International
Battery Materials Association (2014).
Adv. Energy Mater. 2022, 2203153
16146840, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202203153 by Xiamen University, Wiley Online Library on [27/11/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
... Overall, although sulfide-based ASSLSBs are currently in the research and development stage with relatively higher costs, technological advancements and large-scale production will likely lead to a gradual decrease in production costs. Consequently, they are poised to gain a competitive edge over traditional lithium-ion batteries in the future, potentially even replacing them on a large scale, becoming the "main force" in lithium-ion battery technology [67,68]. ...
Article
Full-text available
The sulfide-based all-solid-state lithium-sulfur batteries (ASSLSBs) exhibit superior performance in terms of both safety and energy density compared to the widely used commercial lithium-ion batteries. They hold immense potential for various applications and are expected to drive the development of high-performance energy storage technologies in the future. However, a significant challenge in the practical application of this battery system lies in the (electro)chemical side reactions at the interface between sulfide solid electrolytes (SEs) and lithium an-odes, resulting in interface instability, compromised (electro)chemical performance, and increased safety risks. The above issues seriously hinder the progress of sulfide-based ASSLSBs. This overview delves into the deep-level mechanisms of the interface (electro)chemical reactions, thus thoroughly analyzing the issues associated with the (electro)chemical side reactions at the interface between sulfide SEs and lithium anodes. Moreover, it summarizes the corresponding strategies and latest research advancements in addressing these challenges. Finally, the exploration directions for resolving the (electro)chemical side reactions at the interface between sulfide SEs and lithium anodes, as well as the prospects of sulfide-based ASSLSBs, are presented.
Article
Full-text available
In the ongoing quest to develop lithium‐ion batteries with superior capacity and enhanced safety, the focus has shifted toward all‐solid‐state batteries (SSBs) and nickel‐rich cathode materials. Despite their promise, these technologies face significant interface challenges, notably poor contact and low ion transport efficiency, leading to substantial stability issues. This review aims to provide a comprehensive analysis of both the advantages and the challenges associated with all‐solid‐state batteries. In addition, it discusses the benefits of single‐crystal application in SSBs, in terms of their kinetic performance, mechanical properties, and stability. The review concludes by proposing various strategies to optimize single‐crystal technologies, targeting the development of efficient nickel‐rich single‐crystal materials for use in all‐solid‐state batteries. These approaches offer the potential to address the core challenges currently faced by SSBs and pave the way for the next generation of high‐performance batteries.
Article
All solid-state batteries (ASSBs) represent the next generation of technology, offering tremendous potential in safety and energy density. However, successfully integrating high-capacity cathode materials without compromising long-term stability remains a...
Article
Full-text available
The rapid commercialization of lithium batteries greatly promotes the development of the electric vehicles, renewable energy storage systems and consumer electronics. The service lifetime and safety of lithium batteries are extremely concerned by terminal customers. Sensor technology is powerful in monitoring the physical and chemical signals of lithium batteries, serving for the state of health and safety warning/evaluation of lithium batteries and guide for future development of battery materials. In this review, the primary concern is the generation mechanisms of different physicochemical signals in lithium batteries. Then, the sensing methods and sensing designs according to different physicochemical signals of lithium batteries is summarized. Finally, some perspectives are provided to guide further development of sensing technology for lithium batteries. This review will greatly accelerate the terminal application of advanced sensors, whether in commercial products or scientific research.
Article
Full-text available
Quasi‐solid‐state potassium‐ion batteries (SSPIBs) are of great potential for commercial use due to the abundant reserves and cost‐effectiveness of resources, as well as high safety. Gel polymer electrolytes (GPEs) with high ionic conductivity and fast interfacial charge transport are necessary for SSPIBs. Here, the weak electrostatic force between K⁺ and electronegative functional groups in the ethoxylated trimethylolpropane triacrylate (ETPTA) polymer chains, which can promote fast migration of free K⁺, is revealed. To further enhance the interfacial reaction kinetics, a multilayered GPE by in situ growth of poly(vinylidenefluoride‐co‐hexafluoropropylene) (PVDF‐HFP) on ETPTA (PVDF‐HFP|ETPTA|PVDF‐HFP) is constructed to improve the interface contact and provide sufficient K⁺ concentration in PVDF‐HFP. A high ion transference number (0.92) and a superior ionic conductivity (5.15 × 10⁻³ S cm⁻¹) are achieved. Consequently, the SSPIBs with both intercalation‐type (PB) and conversion‐type (PTCDA) cathodes show the best battery performance among all reported SSPIBs of the same cathode. These findings demonstrate that potassium‐ion batteries have the potential to surpass Li/Na ion batteries in solid‐state systems.
Article
Full-text available
Solid/solid interface is the major challenge for high-performance solid-state batteries. Solid electrolytes (SEs) play a crucial role in the fabrication of effective interfaces in solid-state batteries. Herein, the electrolyte distribution with varied particle sizes is tuned to construct solid-state batteries with excellent performance at different operating temperatures. Solid-state batteries with the configuration S/L (small-sized SE in composite cathode and large-sized SE in electrolyte layer) show the best performance at room temperature (168 mA h g−1 at 0.2 C, retention of 99%, 100 cycles) and −20 °C (89 mA h g−1 at 0.05 C), while the configuration S/S displays better performance at elevated temperatures. The superior performance of the S/L battery is associated with faster lithium-ion dynamics due to the better solid/solid interface between active materials and electrolytes. Moreover, the inferior performance at 60 °C is caused by the formation of voids and cracks in the electrolyte layer during cycling. In contrast, the S/S battery delivers superior performance at elevated operating temperatures because of the integrated structure. This work confirms that tailoring electrolyte size has a significant effect on fabricating all-climate solid-state batteries.
Article
The battery is an indispensable puzzle in the current roadmap towards a carbon-neutral world. With the soaring production, the battery itself is unexpectedly becoming a sustainability problem for society. Increasing attention is thus placed on the life cycle of batteries entailing the second life and recycling of batteries, relying on the monitoring of batteries’ state and the classification of retired batteries by sensors. Battery sensors that decode the fundamental physical/chemical processes are thus poised to maximize the quality, reliability, lifetime, and safety of batteries and to minimize the environmental footprint. Fiber optic sensors stand out by virtue of their miniature size, insulating nature, electromagnetic immunity, and versatile sensitivities. In this Perspective, we discuss the promise and challenges of commercializing smart sensing for batteries and highlight how fiber optic sensors can synergize with paradigm shifts, including cell to pack and cell to chassis technologies.
Article
Inhomogeneity can accelerate performance degradation and reduce the lifetime of large-format batteries and battery modules. It is necessary to in-situ monitor the internal heterogeneous information and investigate the mechanical inhomogeneity of the batteries and modules. In this study, an in-situ measurement platform based on the integrating sensors and optical methods is developed. The internal stress and strain evolutions at the module-level are characterized and related to the deformation evolution at the cell-level. Both levels are suffering the serious mechanical inhomogeneity, which is quantitatively analyzed by the in-situ experiments and corresponding simulations, significantly affecting the long-term cyclic performance and consistency of the module. A new concept of mechanical consistency is introduced to correlate the mechanical inhomogeneity with module aging. This study can provide a basis for analyzing the aging mechanism and optimizing the structure of battery modules.
Article
Enabling lithium metal is key toward maximizing the energy density of solid-state batteries (SSBs) with a solid electrolyte (SE) as the separator. However, the formation of pores at the LiǀSE interface during stripping and short-circuiting during cathodic plating through dendrite growth hinder stable cycling at higher current density. The interface and bulk microstructure of SEs govern the dendrite growth mechanism and hence the maximum or critical current density (CCD) an SE can sustain before shorting. Herein, via the control of the grain size of Li6PS5Cl (LPSCl), we show that interface and bulk microstructure can be simultaneously tailored to significantly influence the CCD. Based on the microstructure-strength relationship, we discuss mechanical failure mechanisms in SSBs derived from correlations between focused ion beam (FIB)-SEM and electrochemical data analyses. We propose that the microstructural influence on local current density and dendrite growth is well connected to the concept of crack growth resistance, or “R curve behavior,” observed in certain materials.
Article
There are abundant electrochemical-mechanical coupled behaviors in lithium-ion battery (LIB) cells on the mesoscale or macroscale level, such as electrode delamination, pore closure, and gas formation. These behaviors are part of the reasons that the excellent performance of LIBs in the lab/material scale fail to transfer to the industrial scale. This paper aims to systematically review these behaviors by utilizing the ‘mechanical origins – structural changes – electrochemical changes – performance’ logic. We first introduce the mechanical origins i.e., the external pressure and internal deformation, based on the different stages of battery life cycle, i.e., manufacture and operation. The response of the batteries due to the two mechanical origins are determined by the mechanical constitutive relation of battery components. The resulting structural changes are ascribed to size and distribution of pores and particles of the battery components, the contact states between different components. The electrochemical changes are divided into ionic/electrical impedance and lifespan. We have summarized massive experimental observations and modelling efforts and the influencing factors in each section. We also clarify the range of external pressure and internal deformation under which the proposed structural and electrochemical changes are likely to take effects. Lastly, we apply the logic to the next generation lithium metal-based solid-state battery. This review will provide useful guidelines to the design and manufacture of lithium-based rechargeable batteries and promote the development of the electric vehicle industry.
Article
Solid-state electrolytes (SSEs) are key to the success and reliability of all-solid-state lithium batteries, potentially enabling improvements in terms of safety and energy density over state-of-the-art lithium-ion batteries. However, there are several critical challenges to their implementation, including the interfacial instability stemming from the dynamic interaction of as-formed dendritic lithium during cycling. For this work, we emphasize the importance of studying the spatial distribution and temporal evolution of strains and defects in crystalline solid-state electrolytes at the micro-scale, and how this affects dendrite growth. A proof-of-principle study is demonstrated using the synchrotron radiation based micro Laue X-ray diffraction method, and a custom-developed in-situ cycling device. Defects and residual strains are mapped, and the evolution of intragranular misorientation is observed. The feasibility of using this technique is discussed, and recommendations for micro-strain engineering to address the Li/SSEs interfacial issues are given. Also, work directions are pointed out with the consideration of combining multi-techniques for “poly-therapy”.
Article
All-solid-state Li metal batteries with ultimate energy density require stabilization at the Li/solid electrolyte interfaces. Further, the use of conventional thick (hundreds of μm) Li metal anodes (LMAs) hinders an impartial evaluation of full cells. In this study, interfacial evolution is compared for thin LMAs (∼ 10 μm thick) prepared by a scalable electrodeposition, with varying protective In/LixIn layer formed via electroless plating. The positive effects of the In/LixIn layer are confirmed for both Li/Li6PS5Cl/Li symmetric cells and LiNi0.70Co0.15Mn0.15O2/Li6PS5Cl/Li full cells. Complementary analysis including electrochemical, ex situ X-ray photoelectron spectroscopy, operando electrochemical pressiometry, and cross-sectional scanning electron microscopy measurements discloses complex Li/Li6PS5Cl interfacial evolutions affected by the In/LixIn coatings, that is, suppressed side reactions and penetrating Li growth. Importantly, a new indicator from operando electrochemical pressiometry analysis, the capacity-normalized pressure change difference Δ(ΔPQ) successfully predicts the dendritic growth of Li during cycling of LiNi0.70Co0.15Mn0.15O2/Li full cell.
Article
Mechanical damages in solid electrolytes of solid-state battery (SSB) during the charging-discharging process remain a challenging issue for battery implementation. This paper demonstrates a numerical simulation of the damages in Li10GeP2S12 (LGPS) solid electrolyte of SSB due to compressive loading generated by electrode volume changes. Three models of anode/electrolyte/cathode arrangements were examined numerically with different expansion-shrinkage behavior. Crack formation inside the electrolyte models was realized by inserting cohesive elements, following traction-separation law. The result shows that when the cathode shrunk and the anode expanded, as occurs in NCM/LGPS/In configuration, the mechanical damages inside the LGPS solid electrolyte are more severe. Due to high-stress generation, there is a plastic deformation in the electrolyte and debonding at the electrode-electrolyte interface. The cracks also appear in both center and edge of the electrolyte because of high-stress concentration. These cracks do not occur when Li4Ti5O12 (LTO) anode with a very low expansion rate is used. This finding confirms that SSB was prone to mechanical damages due to expansion-shrinkage behavior in the electrodes, meaning that the mechanical strength of SSB material constituents must be considered in designing long-lasting SSB.
Article
The influence of external pressure on lithium-ion cells containing a silicon anode is investigated. The performance of pouch-type full cells under rigid compression is compared to unconstrained cells, where the electrode stack is allowed to swell during cycling. The negative electrode contains only silicon as active material, while prelithiated lithium titanium oxide (LTO) is used as the positive electrode. The results show that the main failure mechanism in such cells is a continuous irreversible consumption of lithium ions, likely due to repeated solid electrolyte interphase breakage and reformation. At high pressures, the lithium depletion has a larger influence than at lower pressures. This effect is examined by electrochemical impedance spectroscopy (EIS) and scanning electron microscopy (SEM) as well as dilation measurements of flexibly-constrained cells and can be traced back to an increase of the ionic pore resistance being more pronounced under high pressure. A new approach is used to compensate the lithium loss, i.e. internal relithiation of the LTO electrode via a lithium reservoir. This not only proves the theory of irreversible lithium consumption being the main challenge in these cells, but also enables cycling for 1000 cycles at 1200mAhgSi−1 without capacity fading.
Article
For all-solid-state Li batteries (ASLBs), the external operating pressure offsets the detrimental electrochemo-mechanical effects. In this work, a new scalable in situ protocol to reinforce binders for sulfide-electrolyte-based ASLBs operating under low or no external pressures is reported. The vulcanization of butadiene rubber (BR) using elemental sulfur proceeds in situ during the wet-slurry fabrication process for electrodes, forming a mechanically resilient crosslinked structure. The electrochemical performance of LiNi0.70Co0.15Mn0.15O2 electrodes fabricated using pristine or vulcanized BR diverge significantly as the operating pressure is lowered from 70 MPa to a practically acceptable value of 2 MPa. Complementary analysis using cross-sectional scanning electron microscopy and operando electrochemical pressiometry measurements confirms that the vulcanization of BR suppresses the electrochemo-mechanical degradation of electrodes, which suggests that the scaffolding structure of the vulcanized BR helps maintain the microstructural integrity of the electrodes upon charge and discharge. The significantly enhanced performance of the vulcanized BR is also demonstrated for pouch-type LiNi0.70Co0.15Mn0.15O2/Li4Ti5O12 full cells operated under no external pressure (reversible capacity of 121 vs. 150 mA h g⁻¹ at 0.2C for electrodes with pristine vs. vulcanized BR, respectively).