ArticlePDF Available

Superlattice‐Stabilized WSe2 Cathode for Rechargeable Aluminum Batteries

Authors:

Abstract and Figures

Rechargeable aluminum batteries (RABs), with abundant aluminum reserves, low cost, and high safety, give them outstanding advantages in the postlithium batteries era. However, the high charge density (364 C mm⁻³) and large binding energy of three‐electron‐charge aluminum ions (Al³⁺) de‐intercalation usually lead to irreversible structural deterioration and decayed battery performance. Herein, to mitigate these inherent defects from Al³⁺, an unexplored family of superlattice‐type tungsten selenide‐sodium dodecylbenzene sulfonate (SDBS) (S‐WSe2) cathode in RABs with a stably crystal structure, expanded interlayer, and enhanced Al‐ion diffusion kinetic process is proposed. Benefiting from the unique advantage of superlattice‐type structure, the anionic surfactant SDBS in S‐WSe2 can effectively tune the interlayer spacing of WSe2 with released crystal strain from high‐charge‐density Al³⁺ and achieve impressively long‐term cycle stability (110 mAh g⁻¹ over 1500 cycles at 2.0 A g⁻¹). Meanwhile, the optimized S‐WSe2 cathode with intrinsic negative attraction of SDBS significantly accelerates the Al³⁺ diffusion process with one of the best rate performances (165 mAh g⁻¹ at 2.0 A g⁻¹) in RABs. The findings of this study pave a new direction toward durable and high‐performance electrode materials for RABs.
This content is subject to copyright. Terms and conditions apply.
2201281 (1 of 9) © 2022 Wiley-VCH GmbH
www.small-methods.com
Superlattice-Stabilized WSe2 Cathode for Rechargeable
Aluminum Batteries
Fangyan Cui, Mingshan Han, Wenyuan Zhou, Chen Lai, Yanhui Chen, Jingwen Su,
Jinshu Wang, Hongyi Li,* and Yuxiang Hu*
F. Cui, M. Han, W. Zhou, C. Lai, Y. Chen, J. Su, J. Wang, H. Li, Y. Hu
Key Laboratory of Advanced Functional Materials
Faculty of Materials and Manufacturing
Beijing University of Technology
Beijing 100124, P. R. China
E-mail: lhy06@bjut.edu.cn; y.hu@bjut.edu.cn
DOI: 10.1002/smtd.202201281
density.[4] Amongst, traditional elec-
trodes stability, limited by the inherently
strong Coulombic forces between vulner-
able structures and high-charge-density
aluminum ions (Al3+, 364 C mm3), is the
core challenge in RABs.[5]
To solve above-mentioned issues
in RABs, layered transition metal
dichalcogenides (TMDCs) with moderate
aluminum-selenium/sulfide bonding,
adjustable interlayer spacing, and theo-
retically abundant ions accommodations,
have received much attention in recent
years.[6] Amongst, tungsten selenide (WSe2)
is favorable in metal-ion batteries owing
to its theoretically high capacity, high
conductivity, broad interlayer space, and
moderate metal-selenium bonding interac-
tions, yet never been reported in RABs.[7]
Nevertheless, similar with previous reported
TMDCs in RABs, the strong electrostatic
interaction with inherently high-charge-
density Al3+ (derived from AlxCly in ionic
liquid-based electrolyte) would still result
in irreversible destruction and active materials pulverization/
dissolution (Scheme 1a). Above all, in despite of the advantages
of high theoretical capacity, TMDCs cathodes still suer from
inferior cycling stability. Therefore, the development of a new
strategy to stabilize the WSe2 structure as a case study of TMDCs
is necessary to enhance RABs performance. Superlattice-type
compound, a class of laminated fine materials composed of two
(or more) alternate and periodic components, is especially suit-
able for accommodating high-charge-density active ions.[8] Pre-
viously, superlattice-based materials were proved as stable and
high-performance cathodes in monovalent-ion batteries, such
as LIBs and sodium-ion batteries (SIBs).[9] For example, ordered
superlattice layered oxide (Na3Ni2RuO6) had been applied in
SIBs with two times enhanced capacities in comparison with the
un-treated samples.[8c] The strong Coulombic forces with large
binding energy of multivalent ions make the electrode structure
more prone to irreversible pulverization during cycling than that
of alkali-metal batteries.[10] The superlattice-type materials are pro-
spective candidates for multivalent ions storage due to expanded
interlayer distances, multiple active sites, and improved diusion
kinetics process.[8d] Thus, the introduction of organic molecules
to form superlattice-type structures in TMDCs (such as WSe2)
would be promising to stabilize electrode structure during Al3+
de-intercalation, yet never been explored in RABs.
Rechargeable aluminum batteries (RABs), with abundant aluminum reserves,
low cost, and high safety, give them outstanding advantages in the post-
lithium batteries era. However, the high charge density (364 C mm3) and
large binding energy of three-electron-charge aluminum ions (Al3+) de-
intercalation usually lead to irreversible structural deterioration and decayed
battery performance. Herein, to mitigate these inherent defects from Al3+,
an unexplored family of superlattice-type tungsten selenide-sodium dode-
cylbenzene sulfonate (SDBS) (S-WSe2) cathode in RABs with a stably crystal
structure, expanded interlayer, and enhanced Al-ion diusion kinetic process
is proposed. Benefiting from the unique advantage of superlattice-type struc-
ture, the anionic surfactant SDBS in S-WSe2 can eectively tune the interlayer
spacing of WSe2 with released crystal strain from high-charge-density Al3+
and achieve impressively long-term cycle stability (110 mAh g1 over
1500 cycles at 2.0 A g1). Meanwhile, the optimized S-WSe2 cathode with
intrinsic negative attraction of SDBS significantly accelerates the Al3+
diusion process with one of the best rate performances (165 mAh g1 at
2.0 A g1) in RABs. The findings of this study pave a new direction toward
durable and high-performance electrode materials for RABs.
ReseaRch aRticle
The ORCID identification number(s) for the author(s) of this article
can be found under https://doi.org/10.1002/smtd.202201281.
1. Introduction
The growing demands for state-of-the-art batteries, especially lith-
ium-ion batteries (LIBs) and their large-scale applications raise
the ever-increasing concerns of limited lithium resources, high-
price, and safety issue.[1] In response to these encountered prob-
lems toward LIBs, rechargeable aluminum batteries (RABs) are
promising candidates due to their abundant reserves, low cost,
and high safety.[2] In particular, the aluminum metal anode has
a three-electron-charge transfer and a high volumetric specific
capacity of 8040 mAh cm3 (versus the value of 2046 mAh cm3
for lithium).[3] However, considerable crucial challenges of
RABs still need to be conquered, including electrode pulveri-
zation/solubility, insucient cyclic stability, and low energy
Small Methods 2022, 2201281
23669608, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smtd.202201281 by Henan University, Wiley Online Library on [10/11/2022]. 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
© 2022 Wiley-VCH GmbH
2201281 (2 of 9)
www.advancedsciencenews.com www.small-methods.com
Herein, we proposed a new family of superlattice-type WSe2-
sodium dodecylbenzene sulfonate (SDBS) (S-WSe2) as a model
cathode to overcome the intrinsic weakness in RABs, and sta-
bilize cathodes structure/performance. Benefiting from the
unique superlattice-type structure, the anionic surfactant SDBS
can eectively tune the interlayer spacing of WSe2, release
crystal strain from high-charge-density Al3+, and accelerate the
Al3+ kinetics process, which was further confirmed via theo-
retical calculations. In addition, SDBS as a structural stabilizer
eciently reduce the dissolution of active species in Lewis-acidic
electrolytes, and layered compounds agglomeration. Compared
with traditional WSe2, both superlattice-type and organic inter-
calation strategies could yield enhanced long-term stability.
The optimized S-WSe2 achieved prominent cycling stability
(175 mAh g1 over 500 cycles at 500 mA g1, and above
110 mAh g1 after 1500 cycles at a high current density of
2.0 A g1). Moreover, S-WSe2 cathode exhibited drastically
enhanced rate-capability (165 mAh g1 at 2.0 A g1), and impres-
sively high specific capacity (354 mAh g1 at 100mA g1). Simul-
taneously, in/ex situ characterizations not only revealed the
intercalation reaction mechanism of S-WSe2 but also corrobo-
rated that the facile superlattice-type and organic-intercalated
strategy eciently prevented the structural collapse after long-
term cycling in RABs. Overall, the superlattice-type S-WSe2
cathode would pave a promising direction to design stable and
high-performance electrodes in RABs.
2. Results and Discussions
2.1. Structural Characterization of S-WSe2
Previously, owing to the high-charge-density of Al3+, normal
TMDCs, such as WSe2 usually suered from layered-structure
pulverization and drastic performance reduction in RABs. On
the contrary, the superlattice structure cathodes, which own
inherently stable interfaces, expanded interlayer spacing, and
intercalated organics molecular, drastically stabilize revers-
ible de-intercalation of high-charge-density Al3+, which would
contribute to robust layered-structure and long-term stability
in RABs (Scheme 1b,c). The superlattice-type S-WSe2 was
prepared via a bottom-up one-pot solvothermal approach.
Under the action of a high temperature, high-pressure
transfer physical medium, the raw reactants were constantly
undergoing chemical transport to form new products with con-
trolled particle size, physical phase, and morphology.[11] After
solvothermal reaction at 200 °C for 48 h, the phase composi-
tion and structural information of S-WSe2 and pristine WSe2
are qualitatively identified by X-ray diraction (XRD) patterns
(Figure 1a). The characteristic diraction peaks of WSe2 exhib-
ited at 2θ= 13.42°, 31.29°, 34.32°, and 41.28°, corresponded to
the (002), (100), (102), and (006) crystal planes, respectively,
which were consistent with standard WSe2 phase (JCPDS Card
No. 38–1388).[12] Compared to the WSe2 sample, the XRD pat-
terns of S-WSe2 displayed that the diraction peaks clearly
shifted to a lower angle, with a d-spacing changing from 6.58 Å
in the WSe2 sample to 10.10 Å in the S-WSe2. Moreover, the dif-
fraction peaks of S-WSe2 at 8.74° and 17.03° corresponded to the
(002) and (004) crystal planes, respectively. The obtained results
revealed the intercalation of WSe2 through chain-like organic
molecules (such as sodium dodecylbenzene sulfonate (SDBS))
would eectively and eciently decouple the layers, allowing
for expanding interlayer distance.[13] These features coincided
with the XRD characteristics of the superlattice structure, indi-
cating the successful preparation of superlattice-type S-WSe2.[14]
Furthermore, the Raman spectrum, where the dominant peak
in the S-WSe2 sample appears blue-shifted in comparison with
pristine WSe2 around 250 cm1 in Figure S1 (Supporting Infor-
mation),[15] further confirms the interlayer-expanded S-WSe2.
The content of SDBS in S-WSe2, obtains by thermogravimetric
analysis (TGA) (Figure S2, Supporting Information), is calcu-
lated to be 1.80%. Meanwhile, nitrogen adsorption/desorption
isotherms are conducted (Figure S3, Supporting Information).
The Brunauer–Emmett–Teller (BET) surface area of the S-WSe2
Scheme 1. Schematic evolution of a) traditional WSe2 and b) superlattice-type S-WSe2 electrode after long-term cycling. c) The proposed mechanism
of superlattice-type structure alleviating the pulverization with high-charge-density Al3+ de-intercalation.
Small Methods 2022, 2201281
23669608, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smtd.202201281 by Henan University, Wiley Online Library on [10/11/2022]. 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
© 2022 Wiley-VCH GmbH
2201281 (3 of 9)
www.advancedsciencenews.com www.small-methods.com
(16.44 m2 g1) was higher than the WSe2 sample (12.96 m2 g1),
and the BET value increased by around 25% after the interlay-
expansion treatment. This improvement was mainly ascribed
to the introduction of considerable SDBS, which would also be
beneficial for an enhanced kinetic process.
Furthermore, X-ray photoelectron spectroscopy (XPS) char-
acterization delivers insights into the surface elemental com-
positions and chemical states of superlattice-type S-WSe2,
showing the presence of W 4f, Se 3d, Na 2p, and S 2p peaks
(Figure 1b–d). Double strong and weak peaks are presented
in the 4f orbital of element W (Figure 1b). The strong peaks
belonged to W4+ of S-WSe2 at 32.68eV (W 4f7/2) and 34.63 eV
(W 4f5/2), while the appearance of the weak peaks was ascribed
to the oxidation of W4+ to W6+ (tungsten oxide), in line with
previous literature.[16] The Se 3d XPS spectrum shows the
major doublet peaks of Se2 at 54.50eV (Se 3d5/2) and 55.30eV
(Se 3d3/2) (Figure 1c). Compared with the pristine WSe2, the
presence of Na (Figure 1d) and S elements (Figure S4, Sup-
porting Information) in S-WSe2 originates from the ecacious
introduction of the negative-charged surfactant SDBS. In the
scanning electron microscopy (SEM) (Figure1e) and transmis-
sion electron microscopy (TEM) (Figure1f) images of the super-
lattice-type S-WSe2, there are sparse ultrathin lamellas, while
the morphology of the pure WSe2 exhibits dense nanoflower
patterns (Figure S5, Supporting Information). These results
implied that SDBS organic chained molecules were uniformly
intercalated into the interlayer of WSe2 and further decreased
interlamellar stacking. TEM-energy dispersive spectroscopy
(TEM-EDS, Figure 1i) and SEM-EDS (Figure S6, Supporting
Information) mapping images of S-WSe2 not only reflect that
the related elements exhibit homogeneous distributions but
also demonstrate that S-doping and Na-doping are derived
from electrochemical inertness SDBS. Both the high-resolution
TEM (HR-TEM) image (Figure 1g) and line intensity profiles
of S-WSe2 (Figure1h) results indicate that S-WSe2 layers alter-
nately stack together with the interlayer spacing profoundly
enlarging to 10.1 Å, which match with the XRD analysis. The
emergence of superlattice-type S-WSe2 and interlayer-expanded
structure would reduce the influence from high-charge-density
Al3+ and facilitate aluminum ions diusion with fast reaction
kinetics.[8d,17] Meanwhile, the traditionally TMDCs-optimized
strategy, that was compounding WSe2 with graphene oxide
(GO) (G-WSe2), was also prepared to explore and compare with
the superlattice-type strategy. The Raman, XRD, and TEM char-
acterizations demonstrate the successful introduction of GO
into the WSe2 matrix (Figure S7, Supporting Information), and
GO is partially reduced in G-WSe2 from the ID/IG values calcu-
lated of the Raman spectra.[4d] Typically, the introduction of GO
would alleviate the agglomeration of TMDCs during discharge–
charge cycles. Moreover, the flexible and curled sheet structure
can eectively lessen the volume expansion, thereby to some
extent maintaining the stability of electrode materials.[18]
To further reveal the eect of the amount of SDBS on the
formation of superlattice-type structure, controlled dosages of
SDBS (0, 0.05, 0.10, 0.20, 0.30g) were introduced to synthesize
S-WSe2, named S-WSe2-0, S-WSe2-0.05, S-WSe2-0.10, S-WSe2-
0.20, S-WSe2-0.30. The introduction of the anionic surfactant
SDBS selectively adsorbs onto the (002) crystalline plane of
Figure 1. Characterization of the superlattice-type S-WSe2. a) XRD patterns of S-WSe2 and pristine WSe2. XPS spectra of b) W 4f, c) Se 3d, and d) Na 2p
in S-WSe2. e) SEM image and f) TEM image. g) HR-TEM image of S-WSe2 with h) profile of the interlayer distance in (g). i) HAADF-STEM image and
corresponding elemental mapping of S-WSe2.
Small Methods 2022, 2201281
23669608, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smtd.202201281 by Henan University, Wiley Online Library on [10/11/2022]. 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
© 2022 Wiley-VCH GmbH
2201281 (4 of 9)
www.advancedsciencenews.com www.small-methods.com
WSe2, promoting the growth of the (002) crystalline plane
(Figure S8, Supporting Information). Due to the distinction in
the adsorption capability of SDBS on dierent crystal faces, the
growth rates were variable.[19] Notably, the strong low-angle dif-
fraction peaks appeared only under the optimized addition of
SDBS (0.10 g), indicating the formation of a superlattice-type
structure. Therefore, the S-WSe2-0.10 was optimized as repre-
sentative S-WSe2, while the electrochemical performance was
further conducted.
2.2. Electrochemical Performance of Prepared Samples
The electrochemical performance of WSe2 and S-WSe2 in RABs
was investigated via assembled Swagelok cells. The cyclic vol-
tammetry (CV) curves of the S-WSe2 cathode with voltages
ranging from 0.25V to 1.95V under a scan rate of 0.50mV s1
(Figure 2a). After the formation of solid electrolyte inter-
phase film in the initial cycle,[20] the CV of S-WSe2 exhibited
oxidation and reduction peaks located at around 1.58/1.29
and 1.51/1.20 V, which were attributed to the trivalent Al3+
intercalation/de-intercalation.[21] Moreover, CV tests are car-
ried out at dierent scan rates ranging from 1.0 to 3.0mV s1
(Figure S9, Supporting Information). As scan rates increased,
the CV curves showed similar anodic peak and cathodic
peak. Figure 2b shows the discharge/charge curves of un-
treated WSe2 and superlattice-like S-WSe2 at 100 mA g1 at
room temperature. The initial discharge specific capacity of
S-WSe2 (354 mAh g1) was significantly higher than that of
WSe2 (282 mAh g1). Moreover, the SDBS shows negligible
specific capacity (around 16 mAh g1 at 100 mA g1) in this
system (Figure S10, Supporting Information). The optimized
S-WSe2 with intrinsic negative attraction of SDBS eectively
tuned the interlayer spacing of WSe2 with released crystal
strain from high-charge-density Al3+, and significantly accel-
erated the Al3+ diusion kinetic with enhanced capacity. Alu-
minum ion diusion coecients (D) of WSe2 and S-WSe2
were conducted by the galvanostatic intermittent titration
technique (GITT) test. Compared with the D (1016.4 cm2 s1)
of the WSe2 sample, the S-WSe2 exhibits an order of magni-
tude improvement (1013.7 cm2 s1) with enhanced diusion
kinetic (Figure 2c). The enhanced kinetics process is further
verified via rate capability (Figure 2d). The superlattice-type
S-WSe2 electrode exhibited high capacities of 354, 333, 320,
253, and 165 mAh g1 at current densities of 0.10, 0.20, 0.50,
1.0, and 2.0 A g1, respectively. To the best of our knowledge,
Figure 2. Electrochemical performance of S-WSe2 and WSe2. a) CV curves of S-WSe2 from 0.25 to 1.95V at a scan rate of 0.50mV s1; b) discharge–
charge curves at a current density of 100mA g1; c) galvanostatic intermittent titration technique (GITT) curves; d) comparison of S-WSe2, WSe2, and
other typically reported cathode materials for RABs; e) electrochemical impedance spectroscopy (EIS) curves with fitted curves; f) long-term cycling
stability of S-WSe2 at 500mA g1; g) cycle stability of S-WSe2 at a high current density of 2.0 A g1 and the inset in (g) demonstration of soft-package
RABs lighting LED.
Small Methods 2022, 2201281
23669608, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smtd.202201281 by Henan University, Wiley Online Library on [10/11/2022]. 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
© 2022 Wiley-VCH GmbH
2201281 (5 of 9)
www.advancedsciencenews.com www.small-methods.com
the rate capability of S-WSe2 was among the highest reported
metal-sulfides, small molecules organic, and graphite materials
in RABs, which was mainly ascribed to the certain attraction
to positively charged Al3+ via SDBS and the extended interlayer
for enhanced diusion kinetic.[9d,10a,22] On the contrary, pristine
WSe2 only delivered capacity of 282, 189, 130, and 94 mAh g1 at
current densities of 0.10, 0.20, 0.50, and 1.0 A g1, exhibiting a
much lower rate-performance than that of the superlattice-type
S-WSe2. The improvement of rate-capacity is more pronounced
with increasing current densities (Table S1, Supporting Infor-
mation). The discharge capacity was enhanced by a factor of
around 2.69 (from 94 to 253 mAh g1) at the current density
of 1.0 A g1. Electrochemical impedance spectroscopy (EIS) is
conducted, while experimental results are fitted using an equiv-
alent circuit diagram (Figure S11, Supporting Information) to
clarify the kinetic process of the electrode materials (Figure2e,
Table S2, Supporting Information). The lower Rct of S-WSe2
(298.6 Ω) than the pure sample (5313.0 Ω) played a vital role
in enhanced conductivity and electrochemical reaction kinetics,
which was consistent with GITT test results.[23] The reinforced
diusion and reaction kinetics would provide the possibility for
the application of high-performance RABs.
Compared with the drastical capacity decay of WSe2 after
20 cycles at a current density of 100mA g1 (Figure S12, Supporting
Information), the S-WSe2 retains a high capacity of 175 mAh g1
after 500 cycles at a current density of 500mA g1 (Figure2f ).
Although traditionally TMDC-optimized sample, that is G-WSe2,
exhibits a high initial discharge specific capacity (377 mAh g1)
at 100 mA g1, the retention rate is only 20% after 300 cycles
(77 mAh g1) in Figure S13 (Supporting Information). The inher-
ently unstable architecture of G-WSe2 leads to the collapse of
the structure after the long-term cycling (Figure S14, Supporting
Information). Further, based on XRD pattern and TEM analysis,
G-WSe2 electrode exhibited an amorphous structure after cycles.
The unstable structure leads to its inferior electrochemical
stability (Figure S15, Supporting Information). In particular, the
superlattice-type S-WSe2 could provide a stable crystalline struc-
ture. After 200 cycles, the glass fiber separator of S-WSe2 was still
clean. On the contrary, WSe2 electrode exists a large amount of
irreversible shedding (Figure S16, Supporting Information), and
this may be related to the high compatibility between the unstable
electrode and the Lewis-acidic electrolyte.[24] Furthermore, at
a high current density of 2.0 A g1, the capacity of S-WSe2 still
maintains above 110 mAh g1 even after 1500 cycles (Figure2g),
which is one of the best cycling stability among the high-perfor-
mance TMDCs cathodes in RABs. The inset of Figure2g shows
that the two assembled fully charged pouch cells could lighten
the light emitting diodes (LEDs) and typical discharge–charge
curves of Al/S-WSe2 pouch cells are collected (Figure S17, Sup-
porting Information). The comparison of reported representative
cathode materials in RABs is listed in Table S3 (Supporting
Information), proving the remarkable cycling stability of S-WSe2
cathode with superlattice-type robust structure.
2.3. Mechanism Investigation of Aluminum-storage in S-WSe2
Considering the substantially improved RABs performance of
S-WSe2, it is of great interest to explore the reaction mechanism
in RABs. Thus, a series of in/ex situ characterizations were con-
ducted toward S-WSe2 electrodes. The in situ XRD measure-
ments of the S-WSe2 cathode are conducted via a costumed in
situ testing device (Figure S18, Supporting Information) during
the first discharge–charge cycles (Figure 3a,b). The obtained
diraction peaks matched with the S-WSe2 structure, while
the weak superlattice peaks presented in the range of 8.0–9.0°.
The characteristic peak of the (002) crystal plane slightly shifted
to a lower 2θ value after discharge and backtracked to the initial
state upon charging. This phenomenon could be ascribed
to de-intercalation of active ions, which reflected the reliable
reversibility of the superlattice-type S-WSe2 electrode. More-
over, a shifted and reinstated peak was similar with previous
reported in other metal-ion batteries,[8c,25] which was mainly
ascribed to the robust superlattice-type structure. Meanwhile,
the structural evolution of S-WSe2 was clearly presented in the
enlarged contour maps. For un-treated WSe2, a similar trend of
reversible shift of (002) crystalline plane during the insertion
and extraction of Al-based active ions is observed (Figure S19,
Supporting Information). While, other characteristic peaks
exhibited a certain amorphization during discharge process
and recovered after charge. Consequently, the superlattice-type
S-WSe2 exhibited more stability structure than the pristine
WSe2 under the reaction with high-charge-density Al3+.
Ex situ XPS spectra of the S-WSe2 electrode were further
performed to demonstrate the changes of W, Se, Al, and Cl ele-
ments. The W 4f XPS spectra of the S-WSe2 electrode at the
pristine, fully discharged, and fully re-charged states are pre-
sented in Figure 3c. The main doublet strong peaks of W4+
(32.68 eV (4f7/2) and 34.63 eV (4f5/2)) became weaker and the
new peaks emerged in low binding energy (the emerging
peaks belonging to W2+ (32.13 eV (4f7/2) and 34.23 eV (4f5/2))
and nearing zero-valence W (31.58 eV (4f7/2) and 33.70 eV
(4f5/2), respectively) as the electrode was fully discharged to
0.25 V. The corresponding peaks shifted to low-valance were
ascribed to the partial reduction of tetravalent W (originate
from S-WSe2) and the insertion of aluminum-based active
ions into the S-WSe2.[6e,26] The similar phenomenon exists
for un-treated WSe2 after fully discharged (Figure S20, Sup-
porting Information). After fully charged, the new doublet
peaks of W 4f spectrum in the low binding energy strikingly
decreased, indicating that the low-valance W was oxidized
back to W4+. Note that, there still existed considerable doublet
peaks after fully charged, possibly originating from residual
discharged production. Meanwhile, the XPS spectra of Se 3d
have negligible variation at all three representative states, fur-
ther indicating that the Se element has limited contribution in
the electrochemical reaction (Figure S21, Supporting Informa-
tion). The intensity of Al 2p XPS spectra increases after fully
discharged and then decreases upon re-charging (Figure 3d),
the Cl element has limited variation (Figure S22, Supporting
Information). Meanwhile, there still exists Al 2p peaks in the
S-WSe2 electrode after in-depth XPS etching (Figure S23, Sup-
porting Information), which further reflects that the presence
of Al homogenously exists in the whole S-WSe2 electrode.
More importantly, a real-time transformation of the S-WSe2
cathode was monitored by in situ Raman spectroscopy. The
dominant peak detected at 243 cm1 was ascribed to the W-Se
band of S-WSe2. Besides, other peaks at 349, 310, and 432 cm1
Small Methods 2022, 2201281
23669608, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smtd.202201281 by Henan University, Wiley Online Library on [10/11/2022]. 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
© 2022 Wiley-VCH GmbH
2201281 (6 of 9)
www.advancedsciencenews.com www.small-methods.com
can be assigned to Al-Cl band of AlCl4, Al2Cl7, and Al2Cl7,
respectively (Figure S24, Supporting Information).[2a] During
the discharge process, peaks around 243 cm1 (WSe2 signal of
superlattice-type S-WSe2 electrode) are gradually attenuated
(Figure3e,f ). In the meanwhile, signal of Al2Cl7 decomposed
to AlCl4 and Al3+ (a slight decrease in the stretching vibration
of Al2Cl7 group at 310 and 432 cm1 was accompanied by an
improved intensity of AlCl4 group at 349 cm1), which indi-
cated that the engendered Al3+ participated in the intercalation
process of S-WSe2 cathode and form Alx(S-WSe2). Correspond-
ingly, HR-TEM with EDS was further collected to confirm the
Al3+ insertion and extraction process in the superlattice-type
S-WSe2 cathode. Compared to the discharged stage (Figure3i),
signals of Al present patchy distribution with feeblish den-
sity after fully charged (Figure 3j). These results indicated
that the residual discharged production of S-WSe2 existed
in the charging process, in consistent with the trend in XPS
spectra. Meanwhile, in the contour maps (Figure3g) and three-
dimensional line graphs (Figure3h) of Raman spectra, it can
be observed that WSe2 signals belonging to the S-WSe2 species
are enhanced (the intensity of Raman peaks located in 243 cm1
in the initial re-charged state is low but still exits (Figure S25,
Supporting Information)), namely reversible transformations
between trivalent aluminum ions and S-WSe2. Moreover, we
proposed that the Al-ion storage mechanism in S-WSe2 can be
simplified and formulated as follows
xx
x
()
−++↔
+−
SW
Se Al 3e Al SWSe
2
3
2
(1)
Here the typical aluminum deposition/striping reactions are
delivered by Equation (1). During discharge stage, trivalent Al
ions were inserted into superlattice-type S-WSe2 and transformed
into Alx(S-WSe2), accompanied by the tetravalent tungsten posi-
tive ions undertaking as to the principal redox centers, which
was reversible through the conversion between S-WSe2 and
Alx(S-WSe2) and similar with previous reports in TMDCs.[6e,26]
Moreover, the elemental analysis of discharged S-WSe2 sample
is performed via inductively coupled plasma optical emission
spectrometry (ICP-OES) analysis (Table S4, Supporting Informa-
tion) and TEM-EDS characterization (Figure S26, Supporting
Information). However, due to the complexity of the Lewis-acidic
system in RABs, it was dicult to obtain an accurate x value.[27]
Further in-depth elaborate of the reaction process of TMDCs in
RABs was crucial to the development and application of multiva-
lent-ion batteries, whereas the de-intercalation of aluminum ion,
as verified via a series of in/ex situ characterization, confirmed
superlattice-type S-WSe2 electrode was a promising cathodes in
RABs. Furthermore, the unique long-term reversibility (stability)
of superlattice-type S-WSe2 was investigated in details in the
following part.
2.4. Structural Evolution and Theoretical Simulations
of Superlattice-Type S-WSe2
The superlattice-type materials were a family of promising
electrodes with optimized stability and capacity, which were
Figure 3. Mechanism characterizations of S-WSe2 in RABs. a) In situ XRD patterns of S-WSe2 sample during the initial discharge–charge cycle and
b) the corresponding contour images. Ex situ XPS spectra of c) W 4f and d) Al 2p for S-WSe2 cathode at the pristine, initial fully discharged (0.25V), and
fully charged (1.95V) stages. In situ Raman spectra of S-WSe2 electrode during e,f) the discharge process and g,h) the charge process. HAADF-STEM
image and corresponding elemental mapping images of S-WSe2 electrode after i) fully discharged to 0.25V and j) fully charged to 1.95V.
Small Methods 2022, 2201281
23669608, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smtd.202201281 by Henan University, Wiley Online Library on [10/11/2022]. 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
© 2022 Wiley-VCH GmbH
2201281 (7 of 9)
www.advancedsciencenews.com www.small-methods.com
confirmed and applied in other secondary batteries (alkaline-
metal batteries and two-electron ZIBs).[28] Thus, it was highly
desirable to deeply explore the influence of this property on
three-electron-charge Al-ion. To probe the structural evolution
of superlattice-type S-WSe2 and WSe2 cathode during long-term
cycling, ex situ XRD and TEM characterizations were executed.
Figure 4a exhibits the XRD patterns of S-WSe2 and WSe2 elec-
trodes after long-term cycling (200 cycles). Compared with the
amorphous WSe2 and disappeared characteristic peaks, the
cycled S-WSe2 electrode still retained characteristic diraction
peaks, in compliance with the WSe2 phase. Moreover, density
functional theory (DFT) simulations are conducted to investi-
gate the enhanced Al-storage performance of the superlattice-
like S-WSe2 (Figure4b–d). Simultaneously, TEM images show
that the superlattice-like S-WSe2 electrode retains a comparable
crystal morphology (Figure4f) even after 1000 cycles, while the
WSe2 has become an amorphous structure (Figure 4e) after
200 cycles. The decayed structure of WSe2 would be ascribed
to volume expansion caused by the irreversible transformation
between tetravalent W and divalent W during the charging/
discharging process (volume increased by 86.4%).[29] Moreover,
the deformable material of pristine TMDCs, such as WSe2
was insucient to withstand the insertion/extraction of high-
charge-density Al3+. Moreover, TEM-EDS images reveal that the
W, Se, S, Na, Al, and Cl elements of superlattice-type S-WSe2
cathode still exhibit uniform distribution over 1000 cycles
(Figure S27, Supporting Information), which also demonstrates
the robust structure of superlattice-like S-WSe2 electrode.
On the other hand, DFT simulations were further performed
to explore the energy barrier of superlattice-like S-WSe2 and
untreated WSe2. The electronic structure of S-WSe2 via den-
sity of states (DOS) (Figure S28, Supporting Information),
reflects the spin-up energy band of S-WSe2 crosses the Femi
surface. The S-WSe2 had ignorable bandgap, and its metallic
characteristic delivered high electronic conductivity. Compared
to pristine WSe2, the lower diusion energy barrier of Al atom
diusion in the S-WSe2 is presented (Figure4b). As can be seen
from the models (Figure 4c), aluminum atoms are extremely
easily confined between the upper and lower WSe2 layers.
Whereas, the expanded interlayer of S-WSe2 promotes the alu-
minum atoms to be biased to one side (Figure4d), weakening
the binding force on aluminum and making the de-intercalation
of aluminum ions easier than the original WSe2. Furthermore,
the AlSe bond length of superlattice-type S-WSe2 is longer
(2.71 Å) than that of WSe2 (2.60 Å) and the WSe bond length
has negligible change (Table S5, Supporting Information).
These obtained results signified the intercalation of SDBS has a
slight eect on the intrinsic properties of the active layer, yet sig-
nificantly weakened the binding energy of AlSe bond during
high-charge-density Al3+ intercalation process. The simulations
further confirmed the ultrafast Al storage capability and robust
structure of the superlattice-type S-WSe2 electrode.
3. Conclusion
To overcome the inherent weakness in RABs especially struc-
tural deterioration ascribed to intrinsic high-charge-density
Al3+, we proposed a previously unexplored family of superlat-
tice-type S-WSe2 as a model cathode with drastically improved
long-term stability. The newly introduced anionic organic layer
(typically SDBS) in S-WSe2 not only suppressed crystal strain
with enhanced stability and reduced active-materials dissolu-
tion but also expanded interlayer space with improved diu-
sion process and lower energy barrier, which verified via DFT
simulation. Correspondingly, S-WSe2 exhibited one of the
best cycling stability in TMDCs (110 mAh g1 after 1500 cycles
at a high current density of 2.0 A g1), impressively enhanced
rate performance (165 mAh g1 at 2.0 A g1), and high specific
capacity (354 mAh g1 at 100mA g1). Meanwhile, the reaction
Figure 4. Structural evolution and theoretical simulation of superlattice-type S-WSe2 and WSe2. a) XRD patterns of cycled S-WSe2 and WSe2 electrodes
after 200 cycles. b) Energy profiles of Al diusion on WSe2 and S-WSe2. Typical models of Al diusion path in the interlayer of cycled c) WSe2 and d)
S-WSe2. HR-TEM and the inset corresponding TEM images of e) WSe2 and f) S-WSe2 electrode after long-term cycling tests.
Small Methods 2022, 2201281
23669608, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smtd.202201281 by Henan University, Wiley Online Library on [10/11/2022]. 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
© 2022 Wiley-VCH GmbH
2201281 (8 of 9)
www.advancedsciencenews.com www.small-methods.com
mechanism of superlattice-type S-WSe2 electrode was revealed
in detail via various in/ex situ characterizations. Overall, this
newly superlattice-tuned and organic-intercalated strategy
paves a facile direction to overcome the inherent weakness and
develop high-stability electrodes in RABs.
4. Experimental Section
Materials: Aluminum trichloride (AlCl3, 99.999%) and selenious
acid (H2SeO3, 98%) were obtained from Sigma-Aldrich. 1-ethyl-3-
methylimidazolium chloride (EMICl, 99%) was purchased from
Shanghai Chengjie Chemical Co., Ltd, China. Glass fiber filters (GF/A)
were obtained from Whatman. Aluminum foil (0.25 mm, annealed,
99.99%) was supplied by Alfa Aesar. Sodium tungstate dihydrate
(Na2WO4·2H2O, AR), sodium dodecylbenzene sulfonate (SDBS, AR),
and N,N-Dimethylformamide (DMF, AR) were bought from Sinopharm
Chemical Reagent Co., Ltd, China. Graphene oxide (GO) was obtained
from 3Achem Co., Ltd, China.
Synthesis of WSe2 and G-WSe2: The WSe2 and G-WSe2 were prepared
via the solvothermal synthesis technique. Precisely 1.30g of H2SeO3 and
0.66g of Na2WO4·2H2O were weighed and dissolved in 60 mL of DMF
solution. After sonication and stirring until the mixture was completely
dissolved, then transferred to the polyphenylene reactor, sealed, and
heated at 200 °C for 24h. The obtained product was vacuum filtered and
washed with deionized water and ethanol, and finally dried in a vacuum
oven at 50 °C for 12 h. The G-WSe2 synthesis strategy was similar to
pristine WSe2, besides 10mg GO powders.
Preparation of Superlattice-Type S-WSe2 Nanosheets: The S-WSe2
was synthesized via a reliable bottom-up one-pot solvothermal
approach. Synthesis consists of heating a 5:1 molar ratio of H2SeO3
and Na2WO4·2H2O, and an additional small amount of organic anionic
surfactant SDBS (0.10mg) in a sealed stainless steel reactor at 200 °C
for 48 h to generate the solvothermal product—the S-WSe2 precursor.
The suspension was separated and the final yield of S-WSe2 was stored
in a vacuum drying tank for use.
Characterizations: The morphologies of products were characterized
via SEM (SU8020) and TEM (FEI Talos F200X-G2) characterizations. The
XRD (Bruker, D8 Advance diractometer, Cu Ka, λ= 1.54 Å) patterns of
samples were measured from 5° to 80° (2θ). The nitrogen adsorption-
desorption isotherms were measured via ASAP 2460 and calculated
based on BET analysis. The TGA was performed by a NETZSCH ST
449 F5/F3 Jupiter thermoanalyzer under argon from room temperature to
750 °C with a heating rate of 10 °C min1. The XPS spectra were acquired
through a Thermo Fisher, ESCALAB 250 Xi, and S-WSe2 samples at fully
discharged and fully charged stages were ion-etched using 2000 eV
Argon ion with an etching rate of 0.5 nm s1 under etching time of
25 s. While, Raman spectra were collected in point scan mode using the
Renishaw instrument with a 532nm laser.
Electrochemistry Measurements: All of the WSe2, G-WSe2, and S-WSe2
cathodes were prepared by making a slurry of active material, Ketjen
black, and sodium carboxymethyl cellulose (CMC) in a weight ratio of
6:3:1. Electrochemical measurements were performed using Swagelok
type cells, in which the WSe2, G-WSe2, and S-WSe2 samples were used
as cathodes, high-purity aluminum served as the anode, and Whatman
glass fiber filter acted as separator and conventional Lewis-acid was used
as the electrolyte (AlCl3/EMICl = 1.3:1 = M/M). Galvanostatic charge and
discharge tests were conducted using a Neware battery test station
with a voltage range from 0.25 to 1.95V (vs Al3+/Al). The galvanostatic
intermittent titration technique (GITT) was conducted with pulses
of 6 µA and interruption time for 1.0 h, respectively. Electrochemical
impedance spectroscopy (EIS) was carried out on a PARSTATMC
workstation under frequencies from 0.01 to 105Hz. The CV curves were
measured in a CHI660e electrochemical workstation.
DFT Simulations: The migration barrier energy of Al in WSe2 and
superlattice-type S-WSe2 were calculated using first-principles DFT
under the framework of projector-augmented wave method with
Vienna Ab-initio Simulation Package code. The generalized gradient
approximation of Perdew–Burke–Ernzerh function was used to describe
the exchange and correlation eects. The kinetic energy cuto for plane-
wave expansion was set at 500 eV. A Monkhorst–Pack scheme with
k-spacing of 2π× 0.030 Å1 was used for Brillouin zone sampling. The
convergence criterion for total energy and force was set at 105 eV and
102 eV Å1, respectively.
Supporting Information
Supporting Information is available from the Wiley Online Library or
from the author.
Acknowledgements
This work was financially supported by Beijing Municipal Great Wall
Scholar Training Plan Project (CIT&TCD20190307), Beijing Municipal
Commission of Education (KZ202210005003), National Natural Science
Foundation of China (U1607110 and 51621003), Beijing Natural Science
Foundation (Z210016), Beijing Hundred, Thousand and Ten Thousand
Talent Project (2020016), National Natural Science Foundation of
China (Nos. 52104292, 52130407, and Innovative Research Groups (No.
51621003).
Conflict of Interest
The authors declare no conflict of interest.
Data Availability Statement
The data that support the findings of this study are available from the
corresponding author upon reasonable request.
Keywords
electrode pulverization, long-term stability, rechargeable aluminum
batteries, superlattice-type cathodes, tungsten selenide
Received: October 6, 2022
Published online:
[1] a) J. M. Tarascon, M. Armand, Nature 2001, 414, 359;
b) Y. M. Chiang, Science 2010, 330, 1485; c) B.Dunn, H. Kamath,
J. M. Tarascon, Science 2011, 334, 928; d) W. Hua, X. Yang,
N. P. M.Casati, L.Liu, S.Wang, V.Baran, M.Knapp, H.Ehrenberg,
S.Indris, eScience 2022, 2, 183.
[2] a) M. C. Lin, M. Gong, B. Lu, Y. Wu, D. Y. Wang, M. Guan,
M.Angell, C.Chen, J.Yang, B. J.Hwang, H.Dai, Nature 2015, 520,
324; b) G. A. Elia, K. Marquardt, K. Hoeppner, S.Fantini, R. Lin,
E. Knipping, W. Peters, J. F. Drillet, S. Passerini, R. Hahn, Adv.
Mater. 2016, 28, 7564.
[3] a) Y. Zhang, S. Liu, Y. Ji, J. Ma, H. Yu, Adv. Mater. 2018, 30,
1706310; b) Q. Zhao, M. J. Zachman, W. I. Al Sadat, J. X. Zheng,
L. F.Kourkoutis, L.Archer, Sci. Adv. 2018, 4, eaau8131; c) M. A.Parvez
Mahmud, N.Huda, S. H.Farjana, M.Asadnia, C.Lang, Adv. Energy
Mater. 2017, 8, 1701210; d) X.Zhao, Z. Zhao-Karger, M.Fichtner,
X.Shen, Angew. Chem., Int. Ed. 2019, 59, 5902.
Small Methods 2022, 2201281
23669608, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smtd.202201281 by Henan University, Wiley Online Library on [10/11/2022]. 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
© 2022 Wiley-VCH GmbH
2201281 (9 of 9)
www.advancedsciencenews.com www.small-methods.com
[4] a) T. Gao, X. G. Li, X. W.Wang, J. K. Hu, F. D.Han, X. L. Fan,
L. M. Suo, A. J. Pearse, S. B. Lee, G. W. Rublo, K. J. Gaskell,
M. Noked, C. S. Wang, Angew. Chem., Int. Ed. 2016, 55, 9898;
b) H. C.Yang, H. C.Li, J.Li, Z. H.Sun, K.He, H. M.Cheng, F.Li,
Angew. Chem., Int. Ed. 2019, 58, 11978; c) Y.Hu, D. Sun, B. Luo,
L. Wang, Energy Technol. 2019, 7, 86; d) H. Chen, H. Y. Xu,
S. Y.Wang, T. Q.Huang, J. B.Xi, S. Y.Cai, F.Guo, Z.Xu, W. W.Gao,
C.Gao, Sci. Adv. 2017, 3, eaao7233.
[5] a) Y. Liang, H.Dong, D.Aurbach, Y.Yao, Nat. Energy 2020, 5, 646;
b) T.Koketsu, J.Ma, B. J.Morgan, M.Body, C.Legein, W.Dachraoui,
M. Giannini, A. Demortiere, M. Salanne, F. Dardoize, H. Groult,
O. J.Borkiewicz, K. W.Chapman, P.Strasser, D.Dambournet, Nat.
Mater. 2017, 16, 1142.
[6] a) Z. Hu, X. Liu, P. L. Hernández-Martínez, S. Zhang, P. Gu,
W. Du, W.Xu, H. V. Demir, H. Liu, Q. Xiong, InfoMat 2022, 4,
e12290; b) X.-Y. Yu, L.Yu, X. W. D.Lou, Adv. Energy Mater. 2016, 6,
1501333; c) B.Chen, D. L.Chao, E. Z.Liu, M.Jaroniec, N. Q.Zhao,
S. Z.Qiao, Energy Environ. Sci. 2020, 13, 1096; d) S. Wang, Z.Yu,
J.Tu, J. Wang, D.Tian, Y.Liu, S. Jiao, Adv. Energy Mater. 2016, 6,
1600137; e) S.Guo, H. Yang, M.Liu, X. Feng, H.Xu, Y.Bai, C.Wu,
ACS Appl. Energy Mater. 2021, 4, 7064.
[7] a) P. Wang, F. H. Sun, S. L. Xiong, Z. C. Zhang, B. Duan,
C. H.Zhang, J. K.Feng, B. J.Xi, Angew. Chem., Int. Ed. 2022, 61,
e202116048; b) J.Tang, X.Peng, T.Lin, X.Huang, B.Luo, L.Wang,
eScience 2021, 1, 203.
[8] a) W.Li, C.Han, Q.Gu, S. L.Chou, J. Z.Wang, H. K.Liu, S. X.Dou,
Adv. Energy Mater. 2020, 10, 2001852; b) Y. Jiao, D.Han, Y. Ding,
X.Zhang, G.Guo, J. Hu, D.Yang, A. Dong, Nat. Commun. 2015,
6, 6420; c) Q. Li, S. Xu, S. Guo, K. Jiang, X. Li, M. Jia, P. Wang,
H. Zhou, Adv. Mater. 2020, 32, 1907936; d) P. Xiong, B. Sun,
N. Sakai, R. Ma, T. Sasaki, S. Wang, J. Zhang, G. Wang, Adv.
Mater. 2020, 32, 1902654; e) C. Wang, Q.He, U. Halim, Y. Liu,
E.Zhu, Z.Lin, H. Xiao, X.Duan, Z.Feng, R.Cheng, N. O.Weiss,
G.Ye, Y. C.Huang, H.Wu, H. C.Cheng, I.Shakir, L.Liao, X.Chen,
W. A.GoddardIII, Y.Huang, X.Duan, Nature 2018, 555, 231.
[9] a) J.Li, W.Liu, Z. Yu, J.Deng, S.Zhong, Q. Xiao, F.Chen, D.Yan,
Electrochim. Acta 2021, 370, 137790; b) L.Yao, S.Ju, T. Xu, X.Yu,
ACS Nano 2021, 15, 13662; c) Y. Q.Du, B. Y.Zhang, W. Y.Zhang,
H. X. Jin, J. Y.Qin, J. Q. Wan, J. X. Zhang, G. W. Chen, Energy
Storage Mater. 2021, 38, 231; d) Z. Zhao, Z. Hu, Q. Li, H. Li,
X.Zhang, Y.Zhuang, F.Wang, G.Yu, Nano Today 2020, 32, 100870;
e) W.Bai, J.Gao, K.Li, G.Wang, T. Zhou, P.Li, S.Qin, G.Zhang,
Z.Guo, C.Xiao, Y.Xie, Angew. Chem., Int. Ed. 2020, 59, 17494.
[10] a) Z. J.Lin, M. L.Mao, C. X.Yang, Y. X.Tong, Q. H.Li, J. M.Yue,
G. J.Yang, Q. H.Zhang, L. Hong, X. Q.Yu, L.Gu, Y. S.Hu, H.Li,
X. J. Huang, L. M. Suo, L. Q. Chen, Sci. Adv. 2021, 7, eabg6314;
b) L. Zhou, Q. Liu, Z.Zhang, K.Zhang, F. Xiong, S. Tan, Q.An,
Y. M. Kang, Z. Zhou, L. Mai, Adv. Mater. 2018, 30, 1801984;
c) J. Zhang, Q. Lei, Z. Ren, X. Zhu, J. Li, Z. Li, S. Liu, Y. Ding,
Z.Jiang, J. Li, Y. Huang, X.Li, X.Zhou, Y.Wang, D.Zhu, M. Zeng,
L. Fu, ACS Nano 2021, 15, 17748; d) Y. Shen, Y.Wang, Y. Miao,
M.Yang, X.Zhao, X.Shen, Adv. Mater. 2019, 32, 1905524.
[11] a) M. J.Zhang, Y. D.Duan, C.Yin, M. F.Li, H.Zhong, E.Dooryhee,
K. Xu, F. Pan, F. Wang, J. M. Bai, Sci. Adv. 2020, 6, eabd9472;
b) M.Choucair, P.Thordarson, J. A.Stride, Nat. Nanotechnol. 2009,
4, 30.
[12] H.Jin, Z.Hu, T.Li, L.Huang, J.Wan, G.Xue, J. Zhou, Adv. Funct.
Mater. 2019, 29, 1900649.
[13] a) C. Zhang, B. Fei, D. Yang, H. Zhan, J. Wang, J. Diao, J. Li,
G.Henkelman, D.Cai, J. J.Biendicho, J. R.Morante, A.Cabot, Adv.
Funct. Mater. 2022, 32, 2201322; b) P. A. Zong, D.Yoo, P.Zhang,
Y. Wang, Y.Huang, S.Yin, J. Liang, Y. Wang, K.Koumoto, C.Wan,
Small 2020, 16, e1901901; c) J.Ge, J.Lei, R. N. Zare, Nat. Nano-
technol. 2012, 7, 428.
[14] Z.Wang, R.Li, C.Su, K. P.Loh, SmartMat 2020, 1, e1013.
[15] M. S. Sokolikova, P. C. Sherrell, P. Palczynski, V. L. Bemmer,
C.Mattevi, Nat. Commun. 2019, 10, 712.
[16] B.-Q.Zhang, J.-S.Chen, H.-L.Niu, C.-J.Mao, J.-M.Song, Nanoscale
2018, 10, 20266.
[17] X.Zang, J. N. Hohman, K.Yao, P.Ci, A. Yan, M.Wei, T.Hayasaka,
A. Zettl, P. J.Schuck, J. Wu, L.Lin, Adv. Funct. Mater. 2019, 29,
1807612.
[18] G.Zhang, H. J. Liu, J. H.Qu, J. H.Li, Energy Environ. Sci. 2016, 9,
1190.
[19] a) T.Ito, Crystals 2016, 6, 24; b) A.Szczes, J. Cryst. Growth 2009,
311, 1129; c) L. F.Braganza, M.Dubois, J.Tabony, Nature 1989, 338,
403.
[20] a) S.Zhang, S.Li, Y.Lu, eScience 2021, 1, 163; b) X.Tang, D.Zhou,
B.Zhang, S. J. Wang, P. Li, H.Liu, X. Guo, P.Jaumaux, X. C.Gao,
Y. Z. Fu, C. Y.Wang, C. S.Wang, G. X.Wang, Nat. Commun. 2021,
12, 2857.
[21] a) L. Geng, G. Lv, X. Xing, J. Guo, Chem. Mater. 2015, 27, 4926;
b) B. Jin, S. Hejazi, H. Chu, G. Cha, M. Altomare, M. Yang,
P.Schmuki, Nanoscale 2021, 13, 6087.
[22] a) H. Z. Wang, L. Y. Zhao, H. Zhang, Y. S. Liu, L. Yang, F. Li,
W. H.Liu, X. T.Dong, X. K.Li, Z. H. Li, X. D.Qi, L. Y.Wu, Y. F.Xu,
Y. Q. Wang, K. K.Wang, H. C.Yang, Q.Li, S. S. Yan, X. G.Zhang,
F. Li, H. S. Li, Energy Environ. Sci. 2022, 15, 311; b) D. Y. Wang,
C. Y. Wei, M. C.Lin, C. J. Pan, H. L. Chou, H. A.Chen, M.Gong,
Y. Wu, C. Yuan, M. Angell, Y. J. Hsieh, Y. H. Chen, C. Y. Wen,
C. W.Chen, B. J. Hwang, C. C.Chen, H.Dai, Nat. Commun. 2017,
8, 14283; c) D. J. Yoo, M. Heeney, F.Glocklhofer, J. W. Choi, Nat.
Commun. 2021, 12, 2386; d) D. J. Kim, D.-J. Yoo, M. T. Otley,
A. Prokofjevs, C. Pezzato, M. Owczarek, S. J. Lee, J. W. Choi,
J. F.Stoddart, Nat. Energy 2018, 4, 51.
[23] S.Moon, S. M.Lee, H. K.Lim, H. J.Jin, Y. S.Yun, Adv. Energy Mater.
2021, 11, 2101054.
[24] a) G.-S.Peng, J.Huang, Y.-C.Gu, G.-S.Song, Rare Met. 2021, 40,
3501; b) A. J.Lucio, I.Efimov, O. N.Efimov, C. J.Zaleski, S. Viles,
B. B.Ignatiuk, A. P.Abbott, A. R.Hillman, K. S. Ryder, ChemComm
2021, 57, 9834.
[25] P. Xiong, F.Zhang, X. Zhang, S. Wang, H. Liu, B.Sun, J.Zhang,
Y. Sun, R.Ma, Y.Bando, C.Zhou, Z. Liu, T.Sasaki, G.Wang, Nat.
Commun. 2020, 11, 3297.
[26] X. Q. Feng, J. M. Li, Y. L. Ma, C. Y. Yang, S. Y. Zhang, J. F. Li,
C. S.An, ACS Appl. Energy Mater. 2021, 4, 1575.
[27] E.Faegh, B. Ng, D.Hayman, W. E. Mustain, Nat. Energy 2021, 6,
21.
[28] P. Xiong, Y. Wu, Y. Liu, R. Ma, T.Sasaki, X. Wang, J. Zhu, Energy
Environ. Sci. 2020, 13, 4834.
[29] M.Rahm, R.Homann, N. W.Ashcroft, Chem. - Eur. J. 2016, 22,
14625.
Small Methods 2022, 2201281
23669608, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smtd.202201281 by Henan University, Wiley Online Library on [10/11/2022]. 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
... [48,[69][70] Owing to their unique layered structure, few TMD-based materials, such as MoS 2 , WS 2 , VS 2 , etc., provide a favorable pathway for achieving high storage capacity. [71][72][73][74][75][76][77][78][79][80][81][82][83][84][85] However, TMDs encounter few challenges that restrict their performance and potential in advanced AIBs. Several issues that arise include: i) The limited rate performance and energy density of AIBs can be attributed due to the poor conductivity exhibited by (TMDs). ...
... Apart from MoS 2 and MoSe 2 cathode materials, tungsten disulfide (WS 2 ) and tungsten diselenide (WSe 2 ) have also shown some promising results for AIBs. [79][80][81] For example, Yang et al. fabricated a flexible, free-standing cathode electrode for AIBs. [79] In this work, the authors prepared ultra-small few layered WS 2 nanoplates by nitrogen-doping on the carbon nanofibers (WS 2 @NCNFs) through electrospinning and thermal annealing. ...
... Tungsten dichalcogenide (WSe 2 ), another favorable TMD material, has shown promising prospects as a cathode material for AIBs due to its high theoretical capacity, large interlayer distance, and good electrical conductivity. [81] However, due to the strong columbic electrostatic interaction and high charge density of Al 3 + ions, most of the cathode materials suffer irreversible structural deterioration and poor cycle life. In order Figure 4d shows the charge/discharge curves of star-shaped 2D WS 2 at 0.1 A g À 1 . ...
Article
Full-text available
Rechargeable aluminum‐ion batteries (AIBs) have emerged as a promising candidate for energy storage applications and have been extensively investigated over the past few years. Due to their high theoretical capacity, nature of abundance, and high safety, AIBs can be considered an alternative to lithium‐ion batteries. However, the electrochemical performance of AIBs for large‐scale applications is still limited due to the poor selection of cathode materials. Transition metal dichalcogenides (TMDs) have been regarded as appropriate cathode materials for AIBs due to their wide layer spacing, large surface area, and distinct physiochemical characteristics. This mini‐review provides a succinct summary of recent research progress on TMD‐based cathode materials in non‐aqueous AIBs. The latest developments in the benefits of utilizing 3D‐printed electrodes for AIBs are also explored.
... [8,[10][11] However, the large internal electrostatic force between the inherently high charge density of Al 3+ (364 C mm −3 ) and the fixed host structure results in inferior long-term stability in cathodes, presenting significant scientific challenges in AAIBs. [12] Therefore, the development of suitable cathodes with appropriate diffusion channels is crucial for AAIBs to enable a stable lattice framework and appropriate Coulombic interactions. Prussian blue and its analogs (PBAs) are promising cathodes for multivalent metal-ion batteries owing to their three-dimensional framework and spacious interstitial sites, which effectively facilitate the insertion and extraction of multivalent carriers. ...
Article
Full-text available
Aqueous aluminum ion batteries (AAIBs) hold significant potential for grid‐scale energy storage owing to their intrinsic safety, high theoretical capacity, and abundance of aluminum. However, the strong electrostatic interactions and delayed charge compensation between high‐charge‐density aluminum ions and the fixed lattice in conventional cathodes impede the development of high‐performance AAIBs. To address this issue, this work introduces, for the first time, high‐entropy Prussian blue analogs (HEPBAs) as cathodes in AAIBs with unique lattice tolerance and efficient multipath electron transfer. Benefiting from the intrinsic long‐range disorder and robust lattice strain field, HEPBAs enable the manifestation of the lattice respiration effect and minimize lattice volume changes, thereby achieving one of the best long‐term stabilities (91.2% capacity retention after 10 000 cycles at 5.0 A g⁻¹) in AAIBs. Additionally, the interaction between the diverse metal atoms generates a broadened d‐band and reduced degeneracy compared with conventional Prussian blue and its analogs (PBAs), which enhances the electron transfer efficiency with one of the best rate performance (79.2 mAh g⁻¹ at 5.0 A g⁻¹) in AAIBs. Furthermore, exceptional element selectivity in HEPBAs with unique cocktail effect can facile tune electrochemical behavior. Overall, the newly developed HEPBAs with a high‐entropy effect exhibit promising solutions for advancing AAIBs and multivalent‐ion batteries.
... [23,24] In addition, most chalcogenides exhibit low electron conductivity and a scarcity of active sites due to their wide energy bandgaps. [25,26] Therefore, achieving high stability and enhancing conductivity and reactivity are imperative challenges in constructing high-performance chalcogenide cathodes for AIBs. ...
Article
Full-text available
Chalcogenide cathodes with multi‐electron transfer characteristics are indispensable to aluminum‐ion batteries (AIBs). Nevertheless, their grievous capacity degradation and sluggish reaction kinetics remain fundamental challenges for the practical application. Herein, a Cu2S/Ni3S2 multiphase structure within the metal‐organic frame (MOF) derived carbon decoration layer (CNS@MC) is constructed to elevate the intrinsic electronic properties of chalcogenide cathode and realize high‐performance AIBs. The existence of outer carbon layer and strong orbital interaction at inner heterointerfaces eliminates the bandgap and arises more electrons at Fermi level, efficiently reducing the energy barrier for electron transfer and achieving high reactivity within cathodes. The CNS@MC also presents a strong electronic interaction with active solvent groups, which is beneficial to capture Al³⁺ and facilitate the three‐electron charge‐storage reactions. Experimental results demonstrate that the tailored CNS@MC cathode possesses superior redox kinetics due to the sufficient surface area and rapid Al‐ion diffusion during cycling. Meanwhile, the robust CNS@MC delivers ultra‐high electrochemical stability (131.1 mAh g⁻¹ over 3500 cycles) with high coulombic efficiency and outstanding rate performance. This work offers new opportunities for optimizing the intrinsic properties of the chalcogenide electrodes based on MOF derivatives and heterostructure, providing novel thoughts for designing high‐performance AIBs.
... In addition, the metal-selenium bonds can facilitate the Na + storage kinetics owing to the weaker bond strength than the metal-sulfur bonds. [40,41] However, there are a few reports about the synthesis of such Se-doped V 3 S 4 quantum dots with all the above features in a simple process. ...
Article
Full-text available
Constructing quantum dot‐scale metal sulfides with defects and strongly coupled with carbon is significant for advanced sodium‐ion batteries (SIBs). Herein, Se substituted V3S4 quantum dots with anionic defects confined in nitrogen‐doped carbon matrix (V3S4−xSex/NC) are fabricated. Introducing element Se into V3S4 crystal expands the interlayer distance of V3S4, and triggers anionic defects, which can facilitate Na⁺ diffusions and act as active sites for Na⁺ storage. Meanwhile, the quantum dots tightly encapsulated by conductive carbon framework improve the stability and conductivity of the electrode. Theoretical calculations also unveil that the presence of Se enhances the conductivity and Na⁺ adsorption ability of V3S4−xSex. These properties contribute to the V3S4−xSex/NC with high specific capacity of 447 mAh g⁻¹ at 0.2 A g⁻¹, and prominent rate and cyclic performance with 504 mAh g⁻¹ after 1000 cycles at 10 A g⁻¹. The sodium‐ion hybrid capacitors (SIHCs) with V3S4−xSex/NC anode and activated carbon cathode can achieve high energy/power density (maximum 144 Wh kg⁻¹/5960 W kg⁻¹), capacity retention ratio of 71% after 4000 cycles at 2 A g⁻¹. This work not only synthesizes V3S4−xSex/NC, but also provides a promising opportunity for designing quantum dots and utilizing defects to improve the electrochemical properties.
Article
Despite being promising for high-energy multivalent metal-ion batteries, high-potential and multi-electron-involved Se oxidation conversion usually suffers from rigorous performance decay spawned by sluggish kinetics and active intermediate shuttle/dissolution, especially in...
Article
Full-text available
In the past few decades, great efforts have been made to develop advanced transition metal dichalcogenide (TMD) materials as metal-ion battery electrodes. However, due to existing conversion reactions, they still suffer from structural aggregation and restacking, unsatisfactory cycling reversibility, and limited ion storage dynamics during electrochemical cycling. To address these issues, extensive research has focused on molecular modulation strategies to optimize the physical and chemical properties of TMDs, including phase engineering, defect engineering, interlayer spacing expansion, heteroatom doping, alloy engineering, and bond modulation. A timely summary of these strategies can help deepen the understanding of their basic mechanisms and serve as a reference for future research. This review provides a comprehensive summary of recent advances in molecular modulation strategies for TMDs. A series of challenges and opportunities in the research field are also outlined. The basic mechanisms of different modulation strategies and their specific influences on the electrochemical performance of TMDs are highlighted.
Article
The energy transition to renewables necessitates innovative storage solutions beyond the capacities of lithium‐ion batteries. Aluminum‐ion batteries (AIBs), particularly their aqueous variants (AAIBs), have emerged as potential successors due to their abundant resources, electrochemical advantages, and eco‐friendliness. However, they grapple with achieving their theoretical voltage potential, often yielding less than expected. This perspective article provides a comprehensive examination of the voltage challenges faced by AAIBs, attributing gaps to factors such as the aluminum reduction potential, hydrogen evolution reaction, and aluminum's inherent passivation. Through a critical exploration of methodologies, strategies, such as underpotential deposition, alloying, interface enhancements, tailored electrolyte compositions, and advanced cathode design, are proposed. This piece seeks to guide researchers in harnessing the full potential of AAIBs in the global energy storage landscape.
Article
Full-text available
Rechargeable aluminum batteries (RABs) are an emerging energy storage device owing to the vast Al resources, low cost, and high safety. However, the poor cyclability and inferior reversible capacity of cathode materials have limited the enhancement of RABs performance. Herein, a high configurational entropy strategy is presented to improve the electrochemical properties of RABs for the first time. The high‐entropy (Fe, Mn, Ni, Zn, Mg)3O4 cathode exhibits an ultra‐stable cycling ability (109 mAh g⁻¹ after 3000 cycles), high specific capacity (268 mAh g⁻¹ at 0.5 A g⁻¹), and rapid ion diffusion. Ex situ characterizations indicate that the operational mechanism of (Fe, Mn, Ni, Zn, Mg)3O4 cathode is mainly based on the redox process of Fe, Mn, and Ni. Theoretical calculations demonstrate that the oxygen vacancies make a positive contribution to adjusting the distribution of electronic states, which is crucial for enhancing the reaction kinetics at the electrolyte and cathode interface. These findings not only propose a promising cathode material for RABs, but also provide the first elucidation of the operational mechanism and intrinsic information of high‐entropy electrodes in multivalent ion batteries.
Article
Full-text available
Superlattices are rising stars on the horizon of energy storage and conversion, bringing new functionalities; however, their complex synthesis limits their large‐scale production and application. Herein, a simple solution‐based method is reported to produce organic–inorganic superlattices and demonstrate that the pyrolysis of the organic compound enables tuning their interlayer space. This strategy is exemplified here by combining polyvinyl pyrrolidone (PVP) with WSe2 within PVP/WSe2 superlattices. The annealing of such heterostructures results in N‐doped graphene/WSe2 (NG/WSe2) superlattices with a continuously adjustable interlayer space in the range from 10.4 to 21 Å. Such NG/WSe2 superlattices show a metallic electronic character with outstanding electrical conductivities. Both experimental results and theoretical calculations further demonstrate that these superlattices are excellent sulfur hosts at the cathode of lithium–sulfur batteries (LSB), being able to effectively reduce the lithium polysulfide shuttle effect by dual‐adsorption sites and accelerating the sluggish Li–S reaction kinetics. Consequently, S@NG/WSe2 electrodes enable LSBs characterized by high sulfur usages, superior rate performance, and outstanding cycling stability, even at high sulfur loadings, lean electrolyte conditions, and at the pouch cell level. Overall, this work not only establishes a cost‐effective strategy to produce artificial superlattice materials but also pioneers their application in the field of LSBs.
Article
Full-text available
Van der Waals heterojunctions are fast‐emerging quantum structures fabricated by the controlled stacking of two‐dimensional (2D) materials. Owing to the atomically thin thickness, their carrier properties are not only determined by the host material itself, but also defined by the interlayer interactions, including dielectric environment, charge trapping centers, and stacking angles. The abundant constituents without the limitation of lattice constant matching enable fascinating electrical, optical, and magnetic properties in van der Waals heterojunctions toward next‐generation devices in photonics, optoelectronics, and information sciences. This review focuses on the charge and energy transfer processes and their dynamics in transition metal dichalcogenides (TMDCs), a family of quantum materials with strong excitonic effects and unique valley properties, and other related 2D materials such as graphene and hexagonal‐boron nitride. In the first part, we summarize the ultrafast charge transfer processes in van der Waals heterojunctions, including its experimental evidence and theoretical understanding, the interlayer excitons at the TMDC interfaces, and the hot carrier injection at the graphene/TMDCs interface. In the second part, the energy transfer, including both Förster and Dexter types, are reviewed from both experimental and theoretical perspectives. Finally, we highlight the typical charge and energy transfer applications in photodetectors and summarize the challenges and opportunities for future development in this field. As a fast‐emerging platform, van der Waals heterojunctions have exhibited exotic carrier dynamics in the quantum limit, including charge and energy transfer. Based on the recent experimental and theoretical progress, this review summarizes the state‐of‐art understanding, followed by the representative applications in optoelectronic devices. We also summarize the remaining challenges and opportunities for future development in this field.
Article
Full-text available
Phase separation in conversion/alloying-based anodes easily causes crystal disintegration and leads bad cycling performance. Tin monophosphide (SnP) is an excellent anode material for sodium ion battery due to its unique three-dimensional crystallographic layered structure. In this work, we report the in situ growth of ultrafine SnP nanocrystals within Ti3C2Tx MXene interlayers. The MXene framework is used as a conductive matrix to provide high ionic/electrical transfer paths and reduce the Na⁺ diffusion barrier in the electrode. In situ and ex situ measurements reveal that the synergy between small SnP crystal domains and the confinement provided by the MXene host prevents mechanical disintegration and major phase separation during the sodiation and desodiation cycles. The resultant electrode exhibits fast Na⁺ storage kinetics of 438 mAh g⁻¹ at 15 A g⁻¹ and excellent cycling stability for over 1000 cycles. A full cell assembled with this new SnP-based anode and a Na3V2(PO4)3 cathode delivers a high energy density of 265.4 Wh kg⁻¹ and a power density of 3252.4 W kg–1, outperforming most sodium-ion batteries reported to date.
Article
Full-text available
Lithium-based batteries have had a profound impact on modern society through their extensive use in portable electronic devices, electric vehicles, and energy storage systems. However, battery safety issues such as thermal runaway, fire, and explosion hinder their practical application, especially for using metalanode. These problems are closely related to the high flammability of conventional electrolytes and have prompted the study of flame-retardant and nonflammable electrolytes. Here, we review the recent research on nonflammable electrolytes used in lithium-based batteries, including phosphates, fluorides, fluorinated phosphazenes, ionic liquids, deep eutectic solvents, aqueous electrolytes, and solid-state electrolytes. Their flame-retardant mechanisms and efficiency are discussed, as well as their influence on cell electrochemical performance. We conclude with a summary of future prospects for the design of nonflammable electrolytes and the construction of safer lithium-based batteries.
Article
Full-text available
The practical application of lithium–sulfur batteries is still limited by the lithium polysulfides (LiPSs) shuttling effect on the S cathode and uncontrollable Li‐dendrite growth on the Li anode. Herein, elaborately designed WSe2 flakelets immobilized on N‐doped graphene (WSe2/NG) with abundant active sites are employed to be a dual‐functional host for satisfying both the S cathode and Li anode synchronously. On the S cathode, the WSe2/NG with a strong interaction towards LiPSs can act as a redox accelerator to promote the bidirectional conversion of LiPSs. On the Li anode, the WSe2/NG with excellent lithiophilic features can regulate the uniform Li plating/stripping to mitigate the growth of Li dendrite. Taking advantage of these merits, the assembled Li−S full batteries exhibit remarkable rate performance and stable cycling stability even at a higher sulfur loading of 10.5 mg cm⁻² with a negative to positive electrode capacity (N/P) ratio of 1.4 : 1.
Article
Full-text available
The strong electrostatic interaction between Al ³⁺ and close-packed crystalline structures, and the single-electron transfer ability of traditional cationic redox cathodes, pose challenged for the development of high-performance rechargeable aluminum batteries. Here, to break the confinement of fixed lattice spacing on the diffusion and storage of Al-ion, we developed a previously unexplored family of amorphous anion-rich titanium polysulfides (a-TiS x , x = 2, 3, and 4) (AATPs) with a high concentration of defects and a large number of anionic redox centers. The AATP cathodes, especially a-TiS 4 , achieved a high reversible capacity of 206 mAh/g with a long duration of 1000 cycles. Further, the spectroscopy and molecular dynamics simulations revealed that sulfur anions in the AATP cathodes act as the main redox centers to reach local electroneutrality. Simultaneously, titanium cations serve as the supporting frameworks, undergoing the evolution of coordination numbers in the local structure.
Article
Layered alkali-containing 3d transition-metal oxides are of the utmost importance in the use of electrode materials for advanced energy storage applications such as Li-, Na-, or K-ion batteries. A significant challenge in the field of materials chemistry is understanding the dynamics of the chemical reactions between alkali-free precursors and alkali species during the synthesis of these compounds. In this study, in situ high-resolution synchrotron-based X-ray diffraction was applied to reveal the Li/Na/K-ion insertion-induced structural transformation mechanism during high-temperature solid-state reaction. The in situ diffraction results demonstrate that the chemical reaction pathway strongly depends on the alkali-free precursor type, which is a structural matrix enabling phase transitions. Quantitative phase analysis identifies for the first time the decomposition of lithium sources as the most critical factor for the formation of metastable intermediates or impurities during the entire process of Li-rich layered Li[Li0.2Ni0.2Mn0.6]O2 formation. Since the alkali ions have different ionic radii, Na/K ions tend to be located on prismatic sites in the defective layered structure (Na2/3-x[Ni0.25Mn0.75]O2 or K2/3-x[Ni0.25Mn0.75]O2) during calcination, whereas the Li ions prefer to be localized on the tetrahedral and/or octahedral sites, forming O-type structures.
Article
Rechargeable aluminium-ion batteries (AIBs) are considered to be promising alternatives for current lithium-ion batteries (LIBs), since they can offer the possibilities of low cost with high energy-to-price ratios. Unlike in...