You are on page 1of 14

pubs.acs.

org/JACS Article

Amorphous Chloride Solid Electrolytes with High Li-Ion


Conductivity for Stable Cycling of All-Solid-State High-Nickel
Cathodes
Feng Li,○ Xiaobin Cheng,○ Gongxun Lu, Yi-Chen Yin, Ye-Chao Wu, Ruijun Pan, Jin-Da Luo,
Fanyang Huang, Li-Zhe Feng, Lei-Lei Lu, Tao Ma, Lirong Zheng, Shuhong Jiao, Ruiguo Cao,
Zhi-Pan Liu, Hongmin Zhou, Xinyong Tao,* Cheng Shang,* and Hong-Bin Yao*
Cite This: J. Am. Chem. Soc. 2023, 145, 27774−27787 Read Online
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
Downloaded via CHONNAM NATL UNIV on January 6, 2024 at 09:43:31 (UTC).

ABSTRACT: Solid electrolytes (SEs) are central components that enable


high-performance, all-solid-state lithium batteries (ASSLBs). Amorphous
SEs hold great potential for ASSLBs because their grain-boundary-free
characteristics facilitate intact solid−solid contact and uniform Li-ion
conduction for high-performance cathodes. However, amorphous oxide
SEs with limited ionic conductivities and glassy sulfide SEs with narrow
electrochemical windows cannot sustain high-nickel cathodes. Herein, we
report a class of amorphous Li−Ta−Cl-based chloride SEs possessing high
Li-ion conductivity (up to 7.16 mS cm−1) and low Young’s modulus
(approximately 3 GPa) to enable excellent Li-ion conduction and intact
physical contact among rigid components in ASSLBs. We reveal that the
amorphous Li−Ta−Cl matrix is composed of LiCl43−, LiCl54−, LiCl65−
polyhedra, and TaCl6− octahedra via machine-learning simulation, solid-
state 7Li nuclear magnetic resonance, and X-ray absorption analysis. Attractively, our amorphous chloride SEs exhibit excellent
compatibility with high-nickel cathodes. We demonstrate that ASSLBs comprising amorphous chloride SEs and high-nickel single-
crystal cathodes (LiNi0.88Co0.07Mn0.05O2) exhibit ∼99% capacity retention after 800 cycles at ∼3 C under 1 mA h cm−2 and ∼80%
capacity retention after 75 cycles at 0.2 C under a high areal capacity of 5 mA h cm−2. Most importantly, a stable operation of up to
9800 cycles with a capacity retention of ∼77% at a high rate of 3.4 C can be achieved in a freezing environment of −10 °C. Our
amorphous chloride SEs will pave the way to realize high-performance high-nickel cathodes for high-energy-density ASSLBs.

■ INTRODUCTION
High-energy-density and safe rechargeable batteries are highly
Li9.5Si1.74P1.44S11.7Cl0.3),11,13−15 and revived halides (Li-M-X,
M = non-Li cations, X = halogen anions).16−18 However,
desirable for the ever-increasing demand for sustainable energy incorporating these ceramic SEs in ASSLBs is still hindered by
storage in smart grids and electric vehicles.1−3 All-solid-state battery failure due to rapid interface deterioration19−21 and
lithium batteries (ASSLBs) using nonflammable solid electro- limited ionic conduction percolation in high mass loading
lytes (SEs) instead of liquid electrolytes are promising cathodes (with areal capacity >3 mA h cm−2).22,23
candidates with improved energy/power density, safety, and In contrast to ceramic SEs, amorphous SEs are promising
cycle life.4−6 Although the expected advantages of ASSLBs are candidates for ASSLBs due to their inherent unique glassy
widely acknowledged, their practical applications are still networks for intimate solid−solid contact and excellent
hindered by formidable obstacles stemming from the intrinsic lithium-ion conduction percolation.24−27 Currently, amor-
nature of employed SEs, including low ionic conductivity, phous lithium phosphorus oxynitride (LiPON)28 SEs are
interface incompatibility, and inadequate solid−solid physical being used in commercial thin-film ASSLBs. Nevertheless, the
contact.7,8 energy/power density of LiPON-based thin-film ASSLBs is still
In recent decades, tremendous efforts have been made to
explore suitable SEs for ASSLBs by focusing on oxide, sulfide, Received: September 27, 2023
halide crystalline ceramic, sulfide glassy-ceramic composite, Revised: November 17, 2023
and amorphous oxide and sulfide SEs.5,9−11 Crystalline ceramic Accepted: November 20, 2023
SEs with high ionic conductivities (10−4−10−2 S cm−1) have Published: December 11, 2023
been reported in the oxides (perovskites and garnets),4,12
sulfides (argyrodites, Li6PS5Cl, Li5.5PS4.5Cl1.5 Li10GeP2S12,

© 2023 American Chemical Society https://doi.org/10.1021/jacs.3c10602


27774 J. Am. Chem. Soc. 2023, 145, 27774−27787
Journal of the American Chemical Society pubs.acs.org/JACS Article

inferior to that of current lithium-ion batteries (LIBs) due to synthesis of A-LTC·xLi2O2 (x = 0.25, 0.5, 0.75). Desirable
the low areal capacity of the thin-film cathode and poor ionic stoichiometric amounts of LiCl and TaCl5 powders were mixed and
conductivity (10−6 S cm−1) of LiPON. then loaded into tungsten carbide pots (100 mL) and ball-milled at
To achieve high areal capacity and high voltage cathodes, 450 rpm (6 min of milling, followed by 3 min of rest) in a Planetary
ball mill (YXQM-1L, Changsha Miqi Instruments & Equipment Co.)
highly Li-ion-conductive amorphous SEs with good antiox- for 24−168 h. The mass ratio of tungsten carbide ball mill beads (10,
idative capability are desirable. Presently, ductile sulfide and 8, and 5 mm in diameter; 16, 42, and 52 g in mass) to the precursors
oxysulfide-based amorphous SEs exhibit acceptable Li-ion is 50:1. The pots were opened, and the precursors were ground
conductivities (10−4−10−3 S cm−1).29,30 Unfortunately, it several times to ensure homogenization of the powder during the
remains a great challenge to achieve high areal capacity and milling process. The TaCl5-based composite SEs with other lithium
high voltage cathodes based on oxide, sulfide, or oxysulfide salts and NbCl5-based electrolytes were obtained by introducing
amorphous SEs due to the undesirable trade-off between ionic stoichiometric amounts of Li-X (X = Cl, F, O, OH, O2) in the
conductivity and cathode interfacial compatibility.9,21,31 Aside reaction precursors under the same mixing and ball milling conditions.
from previously explored oxide SEs, chloride SEs have been LiCl and NbCl5 were weighed and mixed in a molar ratio of LiCl/
NbCl5 = 1/1, 4/1 for the synthesis of LiCl-NbCl5 (A-LNC), and
recently proven to overcome the poor oxidative stability of 4LiCl-NbCl5 (A-LNC·3LiCl), respectively. Li2O and NbCl5 were
sulfide SEs.32 Excellent electrochemical performance up to weighed and mixed in a molar ratio of 1/1 Li2O/NbCl5 for the
several thousand cycles has also been demonstrated for bulk- synthesis of Li2O-NbCl5. The obtained fine SE powders were
type ASSLBs with crystalline halides in combination with bare collected and stored for further application and characterization. All
high-voltage cathodes.16−18,33,34 Recently, the ionic conductiv- the procedures were conducted in an argon-filled glovebox (H2O, O2
ity of the crystalline halide is further boosted by the < 0.1 ppm).
modulation of the framework35−37 and the formation of SSW-NN Method. All simulations based on G-NN potential were
glass-ceramic phase38−40 and glass phase,41,42 indicating its conducted by using LASP code45 (Large-scale Atomic Simulation
potential in future application of practical ASSLBs. In addition, with neural network Potential, www.lasphub.com) developed in Liu’s
group. The stochastic surface walking (SSW)46 method was utilized
unique halide-eutectic and clay-like fluoride-halides43,44 had for global optimization, which can automatically explore multidimen-
been revealed with glass phase and excellent ionic conductivity. sional potential energy surfaces (PESs) and identify important
However, their integration into high-areal-capacity ASSLBs is structures and pathways, as demonstrated in our previous work.47,48
hindered by their viscous states. In this context, amorphous The G-NN potential was generated by fitting the DFT global PES
chloride SEs with proper mechanical properties combining the data obtained from the SSW global PES sampling, also known as the
advantages of glass networks and antioxidation chemistry have stochastic surface walking global optimization combined with global
yet to be explored. neural network potential (SSW-NN) method, which typically speeds
Herein, we report a class of amorphous Li−Ta−Cl-based up PES calculations by 3−4 orders of magnitude compared with DFT.
SEs with super Li-ion conductivity up to 7.16 mS cm−1 and This allows for efficient PES exploration at an accuracy comparable
with that of DFT calculations. We generated the triple-element Li−
good compatibility with a high-nickel single-crystal Li- Ta−Cl G-NN potential via the iterative self-learning scheme of the
Ni0.88Co0.07Mn0.05O2 (S-NCM88) cathode. Owing to the soft SSW-NN data set49 by learning LiTaCl6 crystals with atoms ranging
network and high oxidative stability of amorphous Li−Ta−Cl- from 16 to 48. The training data set can be downloaded from (www.
based SEs, the ASSLB with S-NCM88 exhibits a high-rate lasphub.com/supportings/Trainfile_LiTaCl6.tar.gz). The final Li−
capability up to 4 C and good reversibility with a capacity Ta−Cl training data set consists of 5060 structures, as detailed in
retention of 98.8% after 800 cycles at ∼3 C. Also, a well- Table S1. The root-mean-square (rms) errors for the energy and force
operated ASSLB with a high mass loading of S-NCM88 (25 of the G-NN potential are 1.821 meV/atom and 0.027 eV/Å,
mg cm−2) delivering an areal capacity of ∼5 mA h cm−2 is respectively. The important low-energy structures from SSW-NN
calculations have been further examined by using DFT calculations,
demonstrated to retain 80% capacity after 75 cycles at 0.2 C.
and the benchmark results are detailed in Table S2, which shows that
Furthermore, the high-rate capability and life-span of over the energy rms error is below 4 meV/atom for important minima. The
9800 cycles of halide-based ASSLB in a freezing environment accuracy is good enough for the SSW global PES search to find low-
were demonstrated. The present work opens a new avenue for energy minima.
amorphous SE-based ASSLBs with high areal capacity and long DFT Calculations. All DFT calculations were performed by using
cycle life. the plane-wave VASP (Vienna Ab initio Simulation Package) code,50
where the electron−ion interaction is represented by the projector

■ EXPERIMENTAL SECTION
Synthesis of A-LTC and Composite SEs. The starting materials
augmented wave potential. The exchange−correlation function
utilized to train the neural network (NN) potential was the GGA-
PBE,51 and the kinetic energy cutoff was 450 eV. The first Brillouin
used for the synthesis of A-LTC (A represents amorphous, and LTC zone k-point sampling adopted the Monkhorst−Pack scheme with an
represents LiTaCl6) and composite SEs are mixtures of LiCl automated mesh determined by 25 times the reciprocal lattice vectors.
(anhydrous, Aladdin), TaCl5 (anhydrous, Aladdin), and other The energy and force criteria for convergence of the electron density
inorganic lithium salts (LiF, Li2O, LiOH, anhydrous, Aladdin, and structure optimization were set at 5 × 10−6 eV and 0.05 eV/Å,
Li2O2, anhydrous, Sigma-Aldrich). LiCl and TaCl5 were weighed respectively. The dispersion effect is considered by using Grimme’s
and mixed in a molar ratio of LiCl/TaCl5 = 1−7/1 for the synthesis of D3 correction (PBE-D3).52
A-LTC·(x − 1)LiCl (1 ≤ x ≤ 7); LiF, LiCl, and TaCl5 were weighed Li-Ion Conduction Simulation. To identify the equilibrated
and mixed in a molar ratio of LiF/LiCl/TaCl5 = 0.5/1/1, 0.75/1/1, volume (lattice), the initial structure relaxation was always carried out
and 1/1/1 for the synthesis of A-LTC·xLiF (x = 0.5, 0.75, 1); Li2O, for 10 ps using an isothermal−isobaric (NPT) ensemble at 300 K for
LiCl, and TaCl5 were weighed and mixed in a molar ratio of Li2O/ Li−Ta−Cl systems. The Li+ motion was then simulated by the
LiCl/TaCl5 = 0.5/1/1, 1/1/1, and 1.5/1/1 for the synthesis of A- canonical ensemble for A-LTC systems with 512 atoms per supercell
LTC·xLi2O (x = 0.5, 1, 1.5); LiOH, LiCl, and TaCl5 were weighed for 1 ns with a time step of 1 fs, where the first 0.2 ns is for thermal
and mixed in a molar ratio of LiOH/LiCl/TaCl5 = 0.5/1/1, 0.75/1/1, equilibration. To identify the elementary steps of Li-ion migration, we
and 1/1/1 for the synthesis of A-LTC·xLiOH (x = 0.5, 0.75, 1); have optimized the snapshots of the three pathways to local minima
Li2O2, LiCl, and TaCl5 were weighed and mixed in a molar ratio of and located the transition states connecting those minima. The
Li2O2/LiCl/TaCl5 = 0.25/1/1, 0.5/1/1, and 0.75/1/1 for the Einstein relation was used to determine the Li+ diffusion coefficients

27775 https://doi.org/10.1021/jacs.3c10602
J. Am. Chem. Soc. 2023, 145, 27774−27787
Journal of the American Chemical Society pubs.acs.org/JACS Article

(D) from the slopes of the mean square displacements of the Li atom collected by 6 sets of detectors labeled as Bank 2−7 with 2θ range
(Δr2) plotted versus simulation time (t) from 12.5 to 170.0°.56 The diffuse scattering signal originating from
the amorphous component can be clearly observed at low-angle banks
r(t )2 (Bank 2−4) covering a wide d-range, while high-angle banks (Bank
D = lim
t 6t (1) 5−7) possess higher resolution to show more details within a low d-
range.
The activation energies for lithium diffusion (Ea) were computed Solid-State 7Li NMR Measurement. 7Li 1D magic-angle
by fitting the calculated diffusion coefficients to the standard spinning nuclear magnetic resonance (MAS NMR) measurements
Arrhenius form using the following relation were performed on Bruker AVANCE NEO 600 WB spectrometers
i Ea yz (B0 = 14.1 T, Larmor frequency ω0 = 233.24 MHz for 7Li) with a 3.2
D = D0 exp jjj zz mm HXY MAS probe at room temperature. All the powder specimens
k RT { (2) were sealed in ZrO2 tubes with VESPEL caps, and the preparation was
The Ea values obtained were used to characterize the dependence conducted in an Ar-filled glovebox (H2O, O2 < 0.1 ppm) to avoid the
of the averaged diffusion barriers of A-LTC. The conductivity of Li+ degradation of samples induced by moisture. MAS was applied to
ions σ was finally derived from the Nernst−Einstein relation with the obtain high-resolution spectra, with a spinning rate of 22 kHz for 1D
diffusion coefficient D NMR experiments. All 7Li chemical shifts were calibrated with 1 M
LiCl (aq).
q2DN ijj F 2 yzz Electrochemical Measurements of All-Solid-State Batteries.
= j z
VNa jjk RT zz{ (3)
ASSLBs employing amorphous chloride-based SEs in combination
with single-crystal NCM88 (provided by Hefei Gotion High-tech
where Na is the Avogadro constant, q is the charge of the mobile ion Power Energy Co., Ltd.) cathode and a Li−In alloy anode were
(1 for Li+), N is the number of Li atoms, and F is Faraday’s constant. assembled in an argon-filled glovebox (H2O, O2 < 0.1 ppm). S-
Ionic Conductivity and Activation Energy Measurements. NCM88 was mixed with composite SE (A-LTC·Li-X, X = Cl, F, O,
The obtained SE powders were cold-pressed into 10 mm diameter OH, O2, or NbCl5-based electrolytes) and polytetrafluoroethylene
pellets by a hydraulic press (YLJ-15T-LD, Hefei Kejing Materials (PTFE) powders (Guangdong Canrd New Energy Technology Co.,
Technology Co., Ltd.) at 400 MPa for 3 min. Both sides of the pellets Ltd.) at a mass ratio of 70:27:3. The mixtures were then hand-ground
were sputtered with the thin Au layers by a magnetron sputtering for 30 min in an agate mortar and mixed with a miniature vibration
apparatus (SD-900 M, Vision Precision Instruments) for the mixer (MSK-SFM-12 M, Hefei Kejing Materials Technology Co.,
alternating current (AC) impedance measurement. The impedance Ltd.) for another 30 min as the final cathode composite powders.
was measured between −40 and 60 °C with an applied frequency of 1 Li6PS5Cl was employed as a separate layer to construct stable anode-
Hz to 7 MHz and a constant voltage of 50 mV. The electrical electrolyte interface. Approximately 90 mg of SE powder was first
conductivities were investigated by the direct current (DC) placed into a poly(ether ether ketone) (PEEK) cylinder with a 10 mm
polarization method with externally applied voltages ranging from diameter and pressed at 1.5 tons for 2 min to form a separator layer,
100 to 500 mV. All electrochemical measurements were performed and then the cathode composite powder (corresponding to an areal
using a Bio-Logic VMP3 frequency response analyzer in an argon- capacity of 1−5 mA h cm−2) was spread over one side of the separator
filled glovebox (H2O, O2 < 0.1 ppm). A Neware high- and low- layer and pressed at 2 tons for another 2 min. To improve the
temperature test chamber (MGDW-150-40) was applied to control reduction stability of the interface between the SE and anode,
the temperature. approximately 40 mg of Li6PS5Cl (provided by Hefei Gotion High-
XAS Measurement and Analysis. The X-ray absorption fine tech Power Energy Co., Ltd.) powder was dispersed evenly on the
structure spectra (Ta L3-edges) were collected at the 1W1B station at other side of the separator layer and then pressed at 3 tons for 5 min.
the Beijing Synchrotron Radiation Facility (BSRF). The storage rings Then, a piece of In foil (0.1 mm thickness, 10 mm diameter, 3A
of BSRF were operated at 2.5 GeV with an average current of 250 mA. Materials) was attached to the surface of Li6PS5Cl. A Li foil (6 mm
Using a Si(111) double-crystal monochromator, data collection was diameter, China Energy Lithium Co., Ltd.) with a weight ratio of Li/
carried out in transmission/fluorescence mode using an ionization In = 1:50 was subsequently attached to the In foil. Subsequently, the
chamber. All spectra were collected under ambient conditions with ASSLB was pressed at 1 ton for another 1 min and then placed into a
the pellets sealed by Kapton film. The acquired EXAFS data were custom-made stainless-steel casing (Ningbo Zhengli New Energy
processed according to standard procedures using the ATHENA Technology Co., Ltd.) with a constantly applied pressure of 100 MPa
module implemented in the IFEFFIT software package.53−55 The k3- for 24 h to form the Li−In alloy. All of the preparation processes were
weighted EXAFS spectra were obtained by subtracting the post edge conducted in an argon-filled glovebox (H2O, O2 < 0.1 ppm).
background from the overall absorption and then normalizing with Galvanostatic cycling of the ASSLBs was conducted with a constantly
respect to the edge-jump step. Subsequently, k3-weighted χ(k) data of applied pressure of 100 MPa within the voltage range of 2.8−4.3 V
the Ta L3-edge were Fourier transformed to real (R) space using a versus Li+/Li using a LAND-CT2001A (Wuhan, China) and Neware-
hanging window (dk = 1.0 Å−1) to separate the EXAFS contributions CT-4008T battery cycler (Shenzhen, China) at 30 and −10 °C,
from different coordination shells. To obtain the quantitative respectively.
structural parameters around central atoms, least-squares curve Characterizations. Powder X-ray diffraction (PXRD) patterns
parameter fitting was performed by using the ARTEMIS module of were acquired by a Philips X’Pert PRO SUPER X-ray diffractometer
the IFEFFIT software packages. The k3 weighting, k-range of 3.4− using Cu Kα radiation (λ = 1.54178 Å). To prepare the samples, SE
14.2 Å−1, and R range of 1.2−2.7 Å were used for the fitting. The four powders were sealed with a Kapton film in a designed test apparatus,
parameters, coordination number, bond length, Debye−Waller factor, avoiding the effect of ambient humidity. For cryogenic transmission
and E0 shift (CN, R, σ2, ΔE0), were fitted without any being fixed, electron microscopy (cryo-TEM) characterizations, the SE powders
constrained, or correlated. were directly dispersed on a Cu grid and then transferred into the
Neutron Powder Diffraction Measurement and Analysis. cryo-TEM holder in an Ar-filled glovebox (H2O, O2 < 0.1 ppm).
The neutron powder diffraction (NPD) experiment was carried out at Using a sealed container, the cryo-TEM holder was quickly inserted
the multi-physics instrument, a high-intensity time-of-flight diffrac- into the FEI Talos-S TEM chamber, and then liquid nitrogen was
tometer, at the China Neutron Spallation Source (CSNS). The poured into the cryo-TEM holder until the sample temperature
powder sample of about 1.4 g was filled into a cylindrical sample dropped below −170 °C. All cryo-TEM images were taken at
holder and sealed carefully by indium wires under an argon cryogenic temperature (−170 °C). SEM images were obtained by a
atmosphere. Then the container was transferred to the scattering JEOL-JSM-6700F scanning electron microscope with an acceleration
chamber and measured at room temperature with a neutron voltage of 10 kV. Atomic force microscopy (AFM, Dimension Icon,
wavelength band from 0.1 to 4.5 Å. The diffraction patterns were BRUKER) in an argon-filled glovebox (H2O, O2 < 0.1 ppm) was used

27776 https://doi.org/10.1021/jacs.3c10602
J. Am. Chem. Soc. 2023, 145, 27774−27787
Journal of the American Chemical Society pubs.acs.org/JACS Article

Figure 1. Synthesis and Li-ion conductivity evaluation of the obtained A-LTC. (a) PXRD patterns of amorphous LiTaCl6 (A-LTC, top), ball-
milled LiCl (Bm_LiCl, middle) and ball-milled TaCl5 (Bm_TaCl5, bottom), showing the amorphization reaction between LiCl and TaCl5. The
patterns of Bm_LiCl and Bm_TaCl5 are indexed as LiCl (ICSD no. 26909) and TaCl5 (ICSD no. 409433), respectively. (b) Nyquist plots of the
EIS measurement results of A-LTC and Bm_LiCl (inset) at 25 °C. (c) Arrhenius plot of ionic conductivities of the A-LTC together with those of
crystalline sulfide, organic polymer, glassy oxide/sulfide, halide SEs, and LiPF6-based liquid electrolyte with the data adapted with permission
from.13,14,28,29,41,42,44,60−64 The A-LTC possesses a low activation energy and superionic conductivity close to that of the state-of-the-art sulfide SEs.

to analyze the mechanical properties of cold-pressed SE pellets. The precursors revealed that ball-milled LiCl (Bm_LiCl) and
modulus distribution of SE pellets was obtained using Dimension TaCl5 (Bm_TaCl5) under the same conditions were still
Icon (Bruker) in the peak force quantitative nanomechanical mapping maintained in crystalline states (Figure 1a, gray and blue
mode. patterns), which indicates that the mechanochemical reaction
Metal Chlorides Cost Estimation. The comparison of the large-
scale purchase cost of 1000 kg of metal chlorides was based on the between LiCl and TaCl5 is an amorphization process.
method proposed by Hart and Sommerfeld,57 following the following We performed temperature-dependent electrochemical
formula impedance spectroscopy (EIS) measurements to evaluate the
ionic conductivity of the obtained A-LTC. The EIS of A-LTC
log10 P = log10 a + b × log10 Q
displayed high bulk ionic conduction without the grain
In this equation, a and b represent the constants for the given boundary (GB) resistance response (Figure 1b, 25 °C).
materials, P represents the unit price in units of $ kg−1, and Q Representative equivalent circuit model and fitted values are
represents the quantity in units of grams purchased. The prices of the shown in Figure S1a−c and Tables S3, S4. The calculated ionic
metal chlorides with different packing are obtained from the Web site conductivity of A-LTC is 6.05 mS cm−1 (Figure S1a,b) at
of Aladdin (https://www.aladdin-e.com/). room temperature (25 °C). The activation energy for ionic

■ RESULTS AND DISCUSSION


We synthesized amorphous chloride SEs via a mechanochem-
conduction of A-LTC is 0.23 eV based on the ionic
conductivity in the temperature range of −40−60 °C (Figures
1c and S1a−f). In addition, the electronic conductivity of A-
ical reaction of binary compound precursors LiCl and TaCl5 LTC measured by the direct-current (DC) polarization
(Experimental Section). PXRD patterns of the obtained method is ∼5 × 10−9 S cm−1 (Figure S1g,h), implying that
product and precursors are shown in Figure 1a. The featureless A-LTC can be regarded as almost insulating electron transport
pattern (Figure 1a, red pattern) indicates the amorphization of compared to the fast Li-ion conduction. As a control, Bm_LiCl
the precursors to yield an amorphous powder, namely, displayed typical characteristics of poor Li-ion conductors with
amorphous LiTaCl 6 (A-LTC). The peak indexing of a very low ionic conductivity of 6.14 × 10−8 S cm−1 (inset of
27777 https://doi.org/10.1021/jacs.3c10602
J. Am. Chem. Soc. 2023, 145, 27774−27787
Journal of the American Chemical Society pubs.acs.org/JACS Article

Figure 2. SSW-NN simulation and corresponding structure characterizations of the A-LTC. (a) The global PES contour plot of A-LTC systems
containing 3864 distinct low energy minima. The x-axis is a distance-weighted Steinhart OP, the y-axis is the potential energy of A-LTC obtained
from DFT computation, and the color indicates the DOS for the minima. The global minima region is highlighted by a black dashed circle and an
arrow below. (b) Representative local configuration of A-LTC obtained by SSW-NN simulation. (c,d) One-dimensional (1D) 7Li MAS NMR
spectra of the obtained A-LTC (c) and LiCl (d). (e) Corresponding Li-ion proportion of amorphous and crystalline components determined by
the peak integral areas in (c). (f) Fourier transforms of k3-weighted Ta L3-edge EXAFS curves of Ta2O5, TaCl5, and A-LTC. Solid lines and hollow
circles represent raw data and fit results, respectively. (g−j) TEM images with different magnifications (g−i) and the corresponding FFT pattern (j)
of A-LTC.

Figure 1b) and a high activation energy of 0.72 eV (Figure S2). oxide/sulfide,28,29,60,61 halide,41,42,44,63 crystalline sulfide13,14
The apparent difference in Li-ion conduction between A-LTC SEs, and liquid electrolyte64 in Figure 1c. The Li-ion
and Bm_LiCl indicates the important role of TaCl5 in the conductivities and activation energies are also summarized in
formation of the amorphous Li-ion conduction matrix. It Table S5. The superionic conductivity of A-LTC is slightly
seems that all octahedra based on the dimeric structure (− lower than that of the liquid electrolyte and sulfide crystalline
Ta2Cl10−) in pristine TaCl5 are decomposed into separate SEs but outclasses the Li-ion conductivity of polymer, glass-
octahedra (TaCl6−) and interacted with LiCl to sharing ceramic oxide/sulfide, and halide SEs. Furthermore, A-LTC
common edges or corners within the amorphous matrix, which possesses a low activation energy of 0.23 eV, which should
is different from other metal chlorides (ZrCl4, HfCl4, YCl3, endow it with the capability to operate at a wide range of
InCl3, ScCl3 and LaCl3) with maintaining crystalline frame- temperatures and exhibit acceptable Li-ion conductivity at low
works.16−18,58,59 temperatures (0.76 mS cm−1 at −30 °C and 0.44 mS cm−1 at
To compare the Li-ion conductivity of A-LTC with other −40 °C, Figure S1a,c).
SEs, we summarized Arrhenius plots of the ionic conductivity To reveal the atomic arrangement of A-LTC and the
of the A-LTC together with typical polymer,62 glass-ceramic corresponding Li-ion conduction mechanism, we employed
27778 https://doi.org/10.1021/jacs.3c10602
J. Am. Chem. Soc. 2023, 145, 27774−27787
Journal of the American Chemical Society pubs.acs.org/JACS Article

SSW-NN45,46,65 simulation to determine the A-LTC glassy results revealed that A-LTC consists of trace crystalline LiCl
structure. A Li−Ta−Cl NN potential with a fixed component and an amorphous matrix (diffuse peak region), which is
ratio of Li/Ta/Cl = 1:1:6 was first trained based on 5060 consistent with PXRD patterns. The diffuse peaks were
configurations (Table S1) and utilized to determine the obtained by broadening based on long-range ordered peaks.
thermodynamics of LiTaCl6. A total of 30,000 minima are Also, the main simulated peak located at 2.4−3.1 Å
visited with up to 32 atoms in the unit cell, where 3864 distinct corresponds to the main diffuse peak at 2.6−3.3 Å of the
minima in total are collected. Based on those minima, the experimental observation, the corresponding crystallographic
global PES is shown in Figure 2a by plotting the energy against planes were extracted and shown in Figure S6, indicating the
a distance-weighted Steinhart order parameter (OP).66 The reliability of representative local configurations obtained by the
global minimum (GM) phase, which corresponds to the SSW-NN method (detailed discussion can be found in Note
perfect single crystal phase, has a space group P1̅(#2) (Figure S1).
S3a) with the structure composed of LiCl65− and TaCl6− By taking the above representative configurations of the
octahedra sharing a common edge, extending in one dimension amorphous phases as the starting points, we performed 1 ns
while packing stick-by-stick along the other two dimensions. molecular dynamics (MD) simulation for each configuration at
Due to the weak van der Waals interaction between sticks, 300 K to reveal the mechanisms of Li-ion conduction (details
many low-energy minima are energetically degenerated. For can be found in the Experimental Section). From the MD
example, the lowest energy region highlighted by a black trajectories, we have captured three representative migration
dashed circle in Figure 2a corresponds to 11 different pathways. As shown in Figure S7, the energy barriers of the
metastable configurations with subtle differences in their three migration paths of Li-ion are around 0.13 to 0.21 eV,
packing patterns. Such a phenomenon is generally observed where the Li-ion moves from octahedral and tetrahedral voids
in molecular crystals where the packing motif is also mainly to the nearby octahedral and tetrahedral voids. In Movies S1−
determined by the van der Waals interaction.67 Comparatively, S3 we can also see one efficient Li-ion migration that
the most stable phases of ionic crystals generally have distinct contributes to the conductivity may contain several such
local configurations and relatively large energy differences.68 movements. The hopping of Li-ion happens within the
Another important feature is that the highest density of states interspace between Li−Ta−Cl sticks (Figure S7a, Movie S1),
(h-DOS) region in Figure 2a (purple region), which represents within the sticks (Figure S7b, Movie S2), or perpendicular to
the amorphous phase of the material, is rather close to the GM the sticks (Figure S7c, Movie S3). Such a migration
phase. mechanism is similar to that in our recently reported
To show a typical structure of A-LTC, we choose the Li2ZrCl658 with three different migration channels, except
configuration that corresponds to the central point of the h- that the pathways in Li−Ta−Cl involve more intermediates
DOS region in Figure 2a, and the corresponding typical atomic and all energy barriers are lower than that in Li2ZrCl6. The
spatial arrangement consisting of separated vacancies and free computed ionic conductivities of five MD trajectories vary
space is shown in Figures 2b and S3b−e. The amorphous within 2.8−10.4 mS cm−1 and the average conductivity is 6.6
matrix is composed of LiCl43−, LiCl54−, LiCl65− polyhedra, and mS cm−1 at 300 K (Figure S8), which is close to the
TaCl6− octahedra sharing a common edge or corner. All Ta experimental ionic conductivity, identifying the reliability of
atoms remain in the octahedral voids of Cl ions, while the Li the SSW-NN simulation.
atoms are now in the octahedral, rectangular pyramid, and To probe the coordination and local environment of Li
tetrahedral voids. The statistically average pair correlation atoms in A-LTC, we collected the 7Li 1D MAS NMR spectrum
function g(r) and the number of neighboring Cl atoms around of the A-LTC powder. As shown in Figure 2c, the spectrum
the Li and Ta atoms, respectively, of the global PES, are displays two peaks, including a broad peak and a sharp peak
plotted in Figure S4. Both Li atoms and Ta atoms tend to be located at −0.62 and −1.11 ppm, respectively. The presence of
uniformly distributed in amorphous structures, as shown by the peak at −1.11 ppm is recognized from residual LiCl by the
the first peak at 3.57 and 6.24 Å, respectively (Figure S4a,b), comparison of NMR spectroscopy of LiCl powders, as shown
which is consistent with the arrangement model of edge- or in Figure 2d. The assignment of the broad peak at −0.62 ppm
corner-sharing Li−Cl polyhedra and isolated Ta−Cl octahedra to amorphous Li−Ta−Cl was further confirmed by the 7Li
(Figure 2b). Meanwhile, in the amorphous matrix, most Li and NMR spectrum of the Li0.5TaCl5.5 powder by decreasing the
Ta atoms are nearest neighbors with Cl atoms, showing amount of LiCl in the precursors (Figure S9). The broad peak
predominantly by the first peak of Li−Ta, Li−Cl, and Ta−Cl at −0.62 ppm indicates various coordination types of Li with
at 3.69 2.47, and 2.3 Å, respectively, as shown in Figure S4c−e. Cl in the A-LTC that are consistent with the simulated local
Furthermore, the number of neighboring Cl atoms around the configurations (Figure 2b). These two peaks of A-LTC were
Li and Ta atoms is extracted to reveal the coordination mode further deconvoluted with two partially overlapped Lorentz−
of metal cations in the amorphous matrix. In the range of Gaussian peaks (Figure 2c) to quantify the relative proportions
∼2.57−3.87 Å around the Li atoms, the number of of Li atoms in the amorphous Li−Ta−Cl matrix and residual
neighboring Cl atoms fluctuates between 4 and 6 (Figure LiCl, which were determined to be 92.9 and 7.1% (Figure 2e
S4f), which corresponds to the Li−Cl tetrahedron, pentahe- and Table S6), respectively. It is worth noting that residual
dron, and octahedron of the amorphous structure, as shown in LiCl was not detected by PXRD (Figure 1a) owing to the small
Figure 2b. In contrast, the number of Cl atoms around the Ta proportion, as revealed by NMR.
atoms does not fluctuate with distance in the range of ∼2.57− Furthermore, we performed synchrotron X-ray absorption
4.37 Å (Figure S4g), indicating that Ta atoms simply exist as spectroscopy (XAS) to examine the local Ta atomic structure
Ta−Cl octahedra in the amorphous matrix. of the A-LTC matrix. The quite similar Ta L3-edge XANES
Furthermore, NPD was also performed to confirm the spectra of A-LTC, TaCl5, and Ta2O5 (Figure S10) confirm the
consistency of bulk structure between experimental and pentavalent oxidation state of Ta in the obtained A-LTC
simulation of A-LTC. As shown in Figure S5, experimental matrix. Also, the presence of Ta−O signals in A-LTC and
27779 https://doi.org/10.1021/jacs.3c10602
J. Am. Chem. Soc. 2023, 145, 27774−27787
Journal of the American Chemical Society pubs.acs.org/JACS Article

Figure 3. Characterizations of the obtained ALTC·(x − 1)LiCl (2 ≤ x ≤ 7) composite SEs. (a) PXRD patterns of the obtained ALTC·(x − 1)LiCl
(2 ≤ x ≤ 7). The crystalline phase of LiCl is marked with red pentagrams. (b) Ionic conductivity at 25 °C and activation energy of ALTC·(x −
1)LiCl (2 ≤ x ≤ 7). (c) Proposed mosaic model for LiCl-poor and LiCl-rich amorphous matrix-encapsulated nanocrystalline LiCl composite SEs.
The purple irregular geometry, beige rectangle, and purple arrow represent the LiCl nanocrystals, amorphous matrix, and Li-ion migration pathway,
respectively. (d−f) 1D 7Li MAS NMR spectra and deconvolution analysis results corresponding to A-LTC·LiCl (d), A-LTC·3LiCl (e), and A-LTC·
6LiCl (f). (g−k) HRTEM image (g), corresponding FFT result (h) of the selected area (region 1) in (g), magnified HRTEM images (i,j) of the
selected area (regions 2, 3) in (g), and HAADF-STEM combined elemental mapping images (k) of A-LTC·3LiCl. Note: the interface (g) of the
amorphous matrix and nanocrystalline LiCl is selected from the area indicated by the red arrow as presented in (k).

TaCl5 may originate from impurities, including trace Ta2O5 crystalline lattice fringes, which is consistent with the absence
and TaOCl3, generated in the synthesis of A-LTC, and the of diffraction peaks in the PXRD pattern and in Figure 1a, the
transfer process for characterizations. The corresponding red pattern. The amorphous matrix feature of A-LTC can be
extended EXAFS spectra of A-LTC and TaCl5 at the Ta L3- further characterized by the corresponding fast Fourier
edge show an identical dominant peak for the shortest Ta−Cl transform (FFT) pattern with only slight diffraction spots
coordination (Figure 2f and Table S7), revealing the same (Figure 2j), representing the lattice distance of the (111) and
octahedral coordination of Ta (TaCl6−) in both A-LTC and (200) planes of cubic LiCl (referring to ICSD 26909), which is
TaCl5, which is also in agreement with the local TaCl6− consistent with the residual LiCl detected by NMR (Figure
configuration of both distance and coordination numbers 2c). Furthermore, high-angle annular dark-field scanning TEM
obtained by the SSW-NN simulation (Figure S4). (HAADF-STEM) combined with energy-dispersive X-ray
To reveal the microstructure of the A-LTC matrix, we spectroscopy mapping, backscattered electron (BSE) images,
performed cryo-TEM characterizations of the obtained and elemental mapping indicate the homogeneous incorpo-
powders. As shown in Figure 2g, an irregular agglomerate of ration of Ta and Cl into the amorphous matrix (Figure S11).
particles was observed, indicating the random formation of an The amorphous matrix and high polarizability of chlorides
amorphous matrix under mechanochemical reactions. High- also lead to excellent deformability of the synthesized A-LTC.
resolution TEM (HRTEM) images in Figure 2h,i further As shown in Figures S12a−c and S13, intimate contacts
revealed the disordered lattice structure with the absence of without GBs are observed in the cold-pressed A-LTC pellet.
27780 https://doi.org/10.1021/jacs.3c10602
J. Am. Chem. Soc. 2023, 145, 27774−27787
Journal of the American Chemical Society pubs.acs.org/JACS Article

Figure 4. Electrochemical performance characterization of the A-LTC·Li-X (X = F, Cl, O, OH, O2). (a) Optimal ball mill time and ionic
conductivity of A-LTC·Li-X (X = F, Cl, O, OH, O2) electrolytes. (b) LSV curves of A-LTC·Li-X (X = F, Cl, O, OH, O2) electrolyte-based cells.
Testing cell setup: Li−In|Li6PS5Cl|A-LTC·Li-X|A-LTC·Li-X + VGCF. (c) Initial charge/discharge voltage-capacity profiles of ASSLB-Cl, ASSLB-F,
ASSLB-O, AASLB-OH, and ASSLB-O2 operated at current densities of 0.191 mA cm−2 (red profiles) and 1.5 mA cm−2 (blue profiles). (d)
Corresponding mass-normalized dQ/dV curves of ASSLBs. (e) Cycling stability evaluation of ASSLB-Cl, ASSLB-F, ASSLB-O, AASLB-OH, and
ASSLB-O2 at a current density of 1.5 mA cm−2. The filled and hollow symbols represent the capacity and Coulombic efficiency, respectively. The
temperature for all cycling tests is 30 °C.

The smooth surface and excellent ductility of the A-LTC pellet The calculated ionic conductivity and activation energy of A-
endowed by the nature of the amorphous chloride matrix were LTC·(x − 1)LiCl (2 ≤ x ≤ 7) are summarized in Figures 3b,
further confirmed by AFM measurements, showing a low rms S14, and Table S9. All ALTC·(x − 1)LiCl (2 ≤ x ≤ 7) SEs
roughness of 1.66 nm (Figure S12d) and a low average exhibit high ionic conductivity between 1.07 and 7.16 mS cm−1
Young’s modulus of 2.91 ± 0.32 GPa (Figure S12e−i and with a low activation energy of 0.21 ± 0.04 eV. The electronic
Table S8). conductivity of ALTC·(x − 1)LiCl (2 ≤ x ≤ 7) is in the range
To show the superionic conduction and deformability of A- of 4.5 × 10−10−1.0 × 10−9 S cm−1 (Figures S18, S19 and Table
LTC, we prepared a series of A-LTC matrix-encapsulated LiCl S10). It is worth noting that the 7.16 mS cm−1 of room
composite SEs by introducing excessive LiCl in the reaction temperature ionic conductivity of the A-LTC·LiCl composite
precursors. The obtained xLiCl-TaCl5 was denoted as A-LTC· SE is a record-high value reported thus far for advanced
(x − 1)LiCl. The PXRD patterns of the obtained products chloride SEs32 (Figure S20). The high Li-ion conductivity of
reveal the existence of the LiCl phase in the range 2 ≤ x ≤ 7 ALTC·(x − 1)LiCl (2 ≤ x ≤ 7) composite SEs can be
without other crystalline phases (Figure 3a), indicating that attributed to excellent Li-ion migration in the A-LTC matrix
TaCl5 completely reacted with LiCl to form A-LTC. The ionic and at the interface of the matrix/LiCl particles. Specifically, a
conductivity and activation energy of A-LTC·(x − 1)LiCl (2 ≤ moderate amount of LiCl nanocrystals in the amorphous
x ≤ 7) were obtained by temperature-dependent EIS matrix can boost the Li-ion conduction by releasing the mobile
measurements and error bar analysis (Figures S14−S17). Li-ion, which is also observed in the recently reported eutectic
27781 https://doi.org/10.1021/jacs.3c10602
J. Am. Chem. Soc. 2023, 145, 27774−27787
Journal of the American Chemical Society pubs.acs.org/JACS Article

halides.43 Besides, the space charge effect69 widely exists in and Cl elements are distributed uniformly throughout the
composites is beneficial for Li-ion conduction70,71 in the case amorphous matrix, while only the Cl element can be observed
of a high conduction matrix as the crucial factor for overall in the LiCl nanoparticles. The TEM analysis of A-LTC·3LiCl
conductivity of composite SEs. However, with the introduction is consistent with the proposed mosaic model (Figure 3c) that
of more LiCl into the composite SEs, the tortuosity of the Li- the nanocrystalline LiCl particles are dispersed in and around
ion transport pathways increases due to the blocking effect of the A-LTC matrix.
crystalline LiCl (Figure 3c), which results in a decrease in the Using the A-LTC as the matrix, we can also prepare a series
Li-ion conduction in the composite SEs. of composite SEs by encapsulating various lithium salts such as
We further employed 7Li 1D NMR to probe the local LiF, Li2O, LiOH, and Li2O2 (Note S3). We conducted PXRD
environment of Li atoms in the A-LTC·LiCl, A-LTC·3LiCl, measurements (Figure S29a−d) of the obtained composite SEs
and A-LTC·6LiCl composite SEs. As shown in Figure 3d−f and found that all of the samples possess poor crystalline
and Table S6, the resonance chemical shifts of 7Li for the features with tiny crystalline LiCl, indicating that the
amorphous matrix in ALTC·LiCl, ALTC·3LiCl, and ALTC· amorphous matrix possesses good compatibility with different
6LiCl are centered at −0.59, −0.41, and −0.32 ppm, types of anions. Additionally, the Li-ion conduction of the
respectively, which are offset by more positive ppm values obtained composite SEs at room temperature was evaluated by
relative to the A-LTC (−0.62 ppm), indicating the changed EIS, and the Nyquist plots are shown in Figure S29e−h. The
surrounding chemical environments of Li atoms in the calculated room-temperature ionic conductivities are summar-
amorphous matrix. It is assumed that the total number of ized in Figure S29i. All of the A-LTC·Li-X composite SEs
Ta5+ ions in the matrix is constant, whereas the average exhibit good ionic conductivities of 0.78, 4.72, 4.08, and 4.95
number of Ta5+ ions surrounding Li+ ions decreases with the mS cm−1 when appropriate amounts of LiF, Li2O, LiOH, and
introduction of more LiCl, which results in more positive shifts Li2O2 were introduced into the A-LTC matrix, respectively
of the resonance peaks of 7Li. This is consistent with a recent (Figure 4a). These results highlighted the flexibility of A-LTC
result that the more positive resonance chemical shift for 7Li of glassy networks to enable a wide range of composite SE
the halide electrolyte with stacking faults is correlated with a options. However, the different effective ionic radii and
greater number of surrounding Li+ ions.72 Furthermore, the polarizability between halides and oxides resulted in different
insets in Figure 3d−f clearly show that with increasing LiCl deformabilities and solid−solid contact in the composite SEs
precursor content the proportion of residual LiCl increased (Figures S30−S32). It is worth pointing out that the
from 50.9 to 86.1%. introduction of O2−, OH−, and O22− can lead to shorter
Despite the increase in the amount of residual LiCl in the optimal time for achieving the optimal conductivity compared
composite SEs, the deformability of A-LTC is still well to Cl− and F−. As listed in Figure 4a and Table S12, the
maintained. SEM, AFM, and BSE results (Figures S21−S25 optimal conductivities can be achieved in 24 h for A-LTC·Li-X
and Table S8), including the intimate contact GB-free surface, (X = O2−, OH−, and O22−) and over 144 h for A-LTC·Li-X (X
low rms roughness (1.83−2.00 nm), and low average modulus = F−, Cl−), respectively. However, the synthesis of SE for up to
(3.1−3.58 GPa), which is lower than those of oxides and a week would decrease the efficiency and inevitably increase
sulfides (Figure S26), demonstrate the ductility of A-LTC· the time cost of synthesis inevitably. The different synthesis
LiCl, A-LTC·3LiCl, and A-LTC·6LiCl stemming from the soft times may be related to the formation energy of raw materials,
A-LTC matrix. In this scenario, the underlying origin of the as shown in Figure S33. During the reaction, the trend of
exceptional superionic conduction of A-LTC·(x − 1)LiCl (2 ≤ decomposition of raw materials is LiF < LiCl < Li2O < LiOH <
x ≤ 7) composite SEs lies in their intrinsic high ductility and Li2O2. And Ta−O could be more stable than Ta−Cl bonding.
many isotropic migration pathways in the amorphous matrix Thus, when the O2−, OH−, and O22− anions were introduced
(Figure 3c). Besides, the densities of the cold-pressed A-LTC, into the reaction system, the amorphous matrix, consisting of
A-LTC·LiCl, A-LTC·3LiCl, and A-LTC·6LiCl pellets are 3.10, Li−Ta−O−Cl, may form rapidly, resulting in a shorter
2.90, 2.61. 2.23 g cm−3, respectively, which is similar to that of synthesis time compared to that of Cl− and F−.
typical crystalline halides (Figure S27, Note S2, and Table In addition, the effect of various anions on the electro-
S11). chemical oxidation stability of the A-LTC-based composite SE
To reveal the microstructure of the composite SE, we is also investigated by the linear sweep voltammetry (LSV)
performed cryo-TEM on A-LTC·3LiCl. The A-LTC matrix- measurement based on the Li−In|Li6PS5Cl|A-LTC·Li-X|A-
encapsulated LiCl composite structure can be confirmed by LTC·Li-X + VGCF cells. The onset oxidation potentials of A-
selected area electron diffraction (Figure S28). We highlighted LTC, A-LTC·0.5LiF, A-LTC·LiCl, A-LTC·0.5Li2O, A-LTC·
three regions in Figure 3g to show the typical microstructure of 0.75LiOH, and A-LTC·0.5Li2O2 were observed at 4.41, 4.36,
the composite SE. The thick amorphous matrix (Figure 3g, 4.43, 4.30, 4.43, and 4.38 V, respectively (Figure 4b). A similar
region 1) was revealed by the FFT pattern, as shown in Figure oxidation stability may be determined by the characteristic of
3h. At the interface of the amorphous matrix and nanocrystal- amorphous chloride matrix not the various anions, implying
line LiCl (Figure 3g, region 2), distortion of lattice fringes is that all the A-LTC-based composite SEs are compatible with
observed (Figure 3i), which corresponds to the incomplete high-voltage high-nickel cathode (>4 V). Motivated by the
lattice period caused by local strain and defects at the interface. superionic conductivity and desirable electrochemical stability,
The nanocrystalline LiCl (Figure 3g, region 3) was confirmed we fabricated ASSLBs by employing A-LTC-based composite
by the clear lattice with an interplanar spacing of 2.96 Å SEs, S-NCM88 cathodes, and Li−In alloy anodes. Such
(Figure 3j), which matches the crystal plane distance of the ASSLBs can be assembled by a feasible cold-pressing route due
(111) planes of cubic LiCl (referred to as ICSD no. 26909). to the excellent deformability of the A-LTC matrix. The
The LiCl nanoparticles attached on the surface of the A-LTC ASSLBs based on A-LTC·LiCl (7.16 mS cm−1), A-LTC·0.5LiF
aggregate were also revealed by HAADF-STEM and the (0.78 mS cm−1), A-LTC·0.5Li2O (4.72 mS cm−1), A-LTC·
corresponding elemental mapping images (Figure 3k). The Ta 0.75LiOH (4.08 mS cm−1), and A-LTC·0.5Li2O2 (4.95 mS
27782 https://doi.org/10.1021/jacs.3c10602
J. Am. Chem. Soc. 2023, 145, 27774−27787
Journal of the American Chemical Society pubs.acs.org/JACS Article

Figure 5. Performance of the fabricated ASSLB-Cl in the ambient temperature of 30 °C and −10 °C. (a) Long-term cycling at ∼3 C (3 mA cm−2).
(b) Rate performance at various rates in the range from 0.2 to 4.0 C. (c) Cycling characteristics with a high mass loading of S-NCM88 (25.1 mg
cm−2) at 1.0 mA cm−2. (d) Rate performance at various rates in the range from 0.113 (0.25 mA cm−2) to 0.904 C (2 mA cm−2). (e) Discharge
voltage-capacity profiles at various rates in the range from 0.113 (0.25 mA cm−2) to 0.904 C (2 mA cm−2). (f) Discharge capacity retention at
various rates in the range from 0.113 (0.25 mA cm−2) to 0.904 C (2 mA cm−2). (g) Long-term cycling at ∼3.4 C (3 mA cm−2). The ambient
temperature for all cycling tests presented in (a−c,d−g) was 30 and −10 °C, respectively.

cm−1) SEs are denoted as ASSLB-Cl, ASSLB-F, ASSLB-O, reaction. Meanwhile, the corresponding mass-normalized dQ/
AASLB-OH, and ASSLB-O2, respectively. The charge and dV curves also show typical redox peaks to demonstrate the
discharge curves of the initial cycle of these ASSLBs are shown compatibility of A-LTC-based SEs toward S-NCM88 (Figure
in Figure 4c. The initial Coulombic efficiencies (CEs) are 4d). Typical redox reaction peaks are observed at 3.60, 4.01,
94.47, 90.63, 91.13, 91.30, and 94.37% for ASSLB-Cl, ASSLB- and 4.18 V, revealing that the S-NCM88 cathode underwent
F, ASSLB-O, AASLB-OH, and ASSLB-O2, respectively, which three phase transformations, including H1 → M, M → H2, and
indicate that the A-LTC-based composite SEs are stable H2 → H3,73 respectively. In addition, the initial discharge
against the S-NCM88 cathode with a limited interface side capacities of ASSLB-F are 172.48 and 106.15 mA h g−1 at
27783 https://doi.org/10.1021/jacs.3c10602
J. Am. Chem. Soc. 2023, 145, 27774−27787
Journal of the American Chemical Society pubs.acs.org/JACS Article

0.191 and 1.50 mA cm−2, respectively, lower than that of ∼200 indicated effective conduction of Li-ions and intimate physical
and ∼165 mA h g−1 for other ASSLBs (Figure 4c−e), solid−solid contact in the composite cathode enabled by the
indicating lower utilization of S-NCM88 in ASSLB-F. Also, the amorphous matrix even at freezing temperature. To validate
lower utilization is related to the limited percolating network of the high-power capability of ASSLBs at a low temperature,
ion-conduction in the composite cathode resulting from the another ASSLB with an area capacity of 0.88 mA h cm−2 was
loose solid−solid contact and poorer Li-ion conductivity of A- operated at a high current density of up to 3 mA cm−2 (3.37
LTC·0.5LiF (Figure 4a), which led to poorer electrochemical C) for long-term cycling. The ASSLB achieved a capacity
performance. In short, all the fabricated ASSLBs were cycled at retention of over ∼77% over 9000 cycles with an average CE of
a current density of 0.191 mA cm−2 for the first three cycles ∼99.98% (Figure 5g). The corresponding performances are
and the following cycles at 1.5 mA cm−2, all of which exhibited summarized in Table S13 and compared to those of recently
a good capacity retention of >76% and a high average CE of reported ASSLBs based on sulfide and halide SEs. The top-
>99% over 100 cycles (Figure 4e). notch performance of our ASSLB at a low temperature
We further evaluated the long-term cycling stability of demonstrated the potential of amorphous chloride SEs in
ASSLB-Cl due to its best performance among the fabricated tackling the issue of sluggish kinetics of Li-ion migration in the
ASSLBs. As shown in Figure 5a, the ASSLB-Cl can achieve a electrolyte and electrolyte−electrode at low temperatures.
capacity retention of 100% over 800 cycles with an average CE In short, ASSLBs based on A-LTC·LiCl exhibit a high CE,
of ∼98.8%. Another ASSLB-Cl was fabricated and exhibited outstanding rate capability, and excellent cycling stability with
analogical stable operation, in terms of capacity retention of an areal capacity in the range of 1−5 mA h cm−2 in both
83.8 and 70% over 2000 and 3000 cycles (Figure S34), which extreme and conventional environments, which represents the
is the top-notch performance among recent reports. The state-of-the-art performance of recently reported ASSLBs
outstanding cycling performance indicates the realization of based on halide and sulfide SEs, as listed in Figure S37 and
virtually no capacity fading of ASSLB-Cl benefiting from the Tables S13, S14. The amorphous chloride SEs can be
deformability of A-LTC for constructing intimate solid−solid expanded to the cost-effective NbCl5 system, presenting a
contacts and buffering the volume change of the cathode similar capability of Li-ion conduction and a stable operation
during cycling. The rate capability test result is presented in of ASSLBs with an areal capacity of up to 6 mA h cm−2
Figure 5b, and the reversible capacities of ASSLB-Cl are (Figures S38, S39 and Tables S15, S16), detailed discussions
∼205.6, 190.2, 172.2, 149.6, and 132.8 mA h g−1 at rates of can be found in Note S4.
∼0.2, 0.5, 1.0, 2.0, and 3.0 C, respectively. A reversible capacity
of 115.1 mA h g−1 can still be maintained even with a high rate
of ∼4.0 C. The discharge capacity can be recovered when the
■ CONCLUSIONS
In summary, we reported a new type of amorphous chloride Li-
current density returns from 4.0 to 0.2 C. These results ion conductor, Li−Ta−Cl, which can enable the high areal
indicate the excellent rate capability of the fabricated ASSLB- capacity of high-nickel cathodes, owing to its high conductivity,
Cl, which is linked to the excellent electrochemical stability superior mechanical deformability, and electrochemical
and intimate physical contact enabled by the A-LTC matrix. In stability. Based on the A-LTC matrix, various composite SEs
addition, the cycling performance of ASSLB-Cl with a high can be obtained with high ionic conductivity in the range of
mass loading of S-NCM88 (∼25 mg cm−2) is shown in Figure 0.78−7.16 mS cm−1. The fabricated ASSLBs exhibit out-
5c. The reversible phase transitions of S-NCM88 are standing electrochemical performance with a high CE
observable in the dQ/dV curves (Figure S35) at current exceeding 99%, prominent rate capability up to 4 C, and
densities of 0.5 and 1 mA cm−2, indicating adequate Li-ion and excellent capacity retention of 98.8% at 3 C for 800 cycles.
electron conduction in the thick cathode enabled by the Most importantly, the ASSLBs with an areal capacity of up to 5
amorphous matrix. Moreover, ASSLB-Cl displays a remarkable mA h cm−2 exhibited 80% capacity retention after 75 cycles at
reversible areal capacity of 5.1 and 4.9 mA h cm −2 a high current density of 1 mA cm−2. Besides, the high-rate
(corresponding to 205.2 and 195.7 mA h g−1) at 0.5 and 1 capability of ASSLBs in a freezing environment was
mA cm−2, respectively, which provides a competitive areal demonstrated with a stable operation up to 9800 cycles at a
capacity compared to that of commercial LIBs. Under high rate of ∼3.4 C. Our reported amorphous chloride SEs
continuous cycling at 1 mA cm−2, ASSLB-Cl can exhibit a with composition tunability, superionic conductivity, remark-
capacity retention of 80% after 75 cycles and an average CE of able deformability, and good high-nickel cathode compatibility
99.71%. Another two ASSLB-Cl with an S-NCM88 loading of will pave the way to fabricating high-energy-density ASSLBs.
23 mg cm−2, corresponding to an areal capacity of ∼5 mA h
cm−2, also display reproducible performances in terms of
reversible phase transformations, capacity retention, and
■ ASSOCIATED CONTENT
* Supporting Information

average CE, as shown in Figure S36. The Supporting Information is available free of charge at
In order to verify the feasibility of the ASSLBs in a freezing https://pubs.acs.org/doi/10.1021/jacs.3c10602.
environment, we explored the cycling performance of ASSLBs Characterizations, electrochemical results, and additional
at a low temperature of −10 °C. The rate performance of the figures, notes, tables, and movies (PDF)
ASSLB with an area capacity of 2.2 mA h cm−2 is presented in
Crystallographic data of amorphous chloride SEs (ZIP)
Figure 5d. Discharge capacities are 172.5, 162.4, 154.3, 146.2,
128.3, and 114.7 mA h g−1 at current densities of 0.25, 0.5, Crystallographic data of the GM phase material (ZIP)
0.75, 1.0, 1.5, and 2.0 mA cm−2, respectively (Figure 5d,e).
Also, the discharge capacity can be recovered when the current
density returns from 2.0 to 0.5 mA cm−2. Up to 66.5% of the
■ AUTHOR INFORMATION
Corresponding Authors
capacity was retained at a high current density of 2.0 mA cm−2 Xinyong Tao − College of Materials Science and Engineering,
compared to that of 0.25 mA cm−2 (Figure 5f), which Zhejiang University of Technology, Hangzhou 310014
27784 https://doi.org/10.1021/jacs.3c10602
J. Am. Chem. Soc. 2023, 145, 27774−27787
Journal of the American Chemical Society pubs.acs.org/JACS Article

Zhejiang, China; orcid.org/0000-0003-4084-7743; Hefei 230026 Anhui, China; Department of Materials


Email: tao@zjut.edu.cn Science and Engineering, CAS Key Laboratory of Materials
Cheng Shang − Collaborative Innovation Center of Chemistry for Energy Conversion, University of Science and Technology
for Energy Material, Shanghai Key Laboratory of Molecular of China, Hefei 230026 Anhui, China; orcid.org/0000-
Catalysis and Innovative Materials, Key Laboratory of 0003-0860-4151
Computational Physical Science, Department of Chemistry, Ruiguo Cao − Division of Nanomaterials and Chemistry,
Fudan University, Shanghai 200433, China; Shanghai Qi Hefei National Research Center for Physical Sciences at the
Zhi Institute, Shanghai 200030, China; orcid.org/0000- Microscale, University of Science and Technology of China,
0001-7486-1514; Email: cshang@fudan.edu.cn Hefei 230026 Anhui, China; Department of Materials
Hong-Bin Yao − Division of Nanomaterials and Chemistry, Science and Engineering, CAS Key Laboratory of Materials
Hefei National Research Center for Physical Sciences at the for Energy Conversion, University of Science and Technology
Microscale, University of Science and Technology of China, of China, Hefei 230026 Anhui, China; orcid.org/0000-
Hefei 230026 Anhui, China; Department of Applied 0002-3177-3917
Chemistry, University of Science and Technology of China, Zhi-Pan Liu − Collaborative Innovation Center of Chemistry
Hefei 230026 Anhui, China; orcid.org/0000-0002-2901- for Energy Material, Shanghai Key Laboratory of Molecular
0160; Email: yhb@ustc.edu.cn Catalysis and Innovative Materials, Key Laboratory of
Computational Physical Science, Department of Chemistry,
Authors Fudan University, Shanghai 200433, China; Shanghai Qi
Feng Li − Division of Nanomaterials and Chemistry, Hefei Zhi Institute, Shanghai 200030, China; orcid.org/0000-
National Research Center for Physical Sciences at the 0002-2906-5217
Microscale, University of Science and Technology of China, Hongmin Zhou − Division of Nanomaterials and Chemistry,
Hefei 230026 Anhui, China Hefei National Research Center for Physical Sciences at the
Xiaobin Cheng − Department of Applied Chemistry, Microscale, University of Science and Technology of China,
University of Science and Technology of China, Hefei 230026 Hefei 230026 Anhui, China
Anhui, China Complete contact information is available at:
Gongxun Lu − College of Materials Science and Engineering, https://pubs.acs.org/10.1021/jacs.3c10602
Zhejiang University of Technology, Hangzhou 310014
Zhejiang, China Author Contributions
Yi-Chen Yin − Department of Applied Chemistry, University ○
F.L. and X.C. contributed equally to this work.
of Science and Technology of China, Hefei 230026 Anhui, Notes
China The authors declare no competing financial interest.
Ye-Chao Wu − Department of Applied Chemistry, University
of Science and Technology of China, Hefei 230026 Anhui,
China; Hefei Gotion High-tech Power Energy Co., Ltd., Hefei
230012 Anhui, China
■ ACKNOWLEDGMENTS
We acknowledge the financial support from the National
Ruijun Pan − Hefei Gotion High-tech Power Energy Co., Ltd., Natural Science Foundation of China (grant nos. 22325505,
Hefei 230012 Anhui, China 52073271, 52225208, 22161142004, 22033003, 92061112,
Jin-Da Luo − Department of Applied Chemistry, University of and 22122301), the National Key Research and Development
Science and Technology of China, Hefei 230026 Anhui, Program of China (grant no. 2018YFA0208600), the
China Fundamental Research Funds from University of Science and
Fanyang Huang − Division of Nanomaterials and Chemistry, Technology of China (grant no. KY2060000207), the
Hefei National Research Center for Physical Sciences at the Collaborative Innovation Program of Hefei Science Center,
Microscale, University of Science and Technology of China, CAS (grant no. 2022HSC-CIP018), and the Open Funds of
Hefei 230026 Anhui, China; Department of Materials the State Key Laboratory of Rare Earth Resource Utilization
Science and Engineering, CAS Key Laboratory of Materials (grant no. RERU2022003). T.M. acknowledges the Shenzhen
for Energy Conversion, University of Science and Technology Science and Technology Program (grant no.
of China, Hefei 230026 Anhui, China JCYJ20190808143007479). We thank the beamline 1W1B
Li-Zhe Feng − Department of Applied Chemistry, University station of Beijing Synchrotron Radiation Facility (BSRF) for
of Science and Technology of China, Hefei 230026 Anhui, providing the beam time. We greatly appreciate the neutron
China beamtime granted from the CSNS and the technical assistance
Lei-Lei Lu − Division of Nanomaterials and Chemistry, Hefei from Juping Xu, Wen Yin, Zhongyuan Huang and Yinguo
National Research Center for Physical Sciences at the Xiao. We thank the support from the USTC Center for Micro-
Microscale, University of Science and Technology of China, and Nanoscale Research and Fabrication. We also thank the
Hefei 230026 Anhui, China support from the Hefei Advanced Computing Center. We are
Tao Ma − Division of Nanomaterials and Chemistry, Hefei grateful to Dr. L.Z. from the Institute of High Energy Physics
National Research Center for Physical Sciences at the for the analysis of EXAFS results. We are also grateful to Dr.
Microscale, University of Science and Technology of China, Ke Gong and Dr. Yu-Song Wang from the USTC Center for
Hefei 230026 Anhui, China Physical and Chemical Science Laboratory for the analysis of
Lirong Zheng − Institute of High Energy Physics, the Chinese NMR spectra.
Academy of Sciences, Beijing 100049, China
Shuhong Jiao − Division of Nanomaterials and Chemistry,
Hefei National Research Center for Physical Sciences at the
■ REFERENCES
(1) Dunn, B.; Kamath, H.; Tarascon, J.-M. Electrical energy storage
Microscale, University of Science and Technology of China, for the grid: a battery of choices. Science 2011, 334 (6058), 928−935.

27785 https://doi.org/10.1021/jacs.3c10602
J. Am. Chem. Soc. 2023, 145, 27774−27787
Journal of the American Chemical Society pubs.acs.org/JACS Article

(2) Lin, D.; Liu, Y.; Cui, Y. Reviving the lithium metal anode for (23) Randau, S.; Weber, D. A.; Kötz, O.; Koerver, R.; Braun, P.;
high-energy batteries. Nat. Nanotechnol. 2017, 12 (3), 194−206. Weber, A.; Ivers-Tiffée, E.; Adermann, T.; Kulisch, J.; Zeier, W. G.;
(3) Larcher, D.; Tarascon, J.-M. Towards greener and more et al. Benchmarking the performance of all-solid-state lithium
sustainable batteries for electrical energy storage. Nat. Chem. 2015, batteries. Nat. Energy 2020, 5 (3), 259−270.
7 (1), 19−29. (24) Minami, T. Fast ion conducting glasses. J. Non Cryst. Solids
(4) Han, X.; Gong, Y.; Fu, K. K.; He, X.; Hitz, G. T.; Dai, J.; Pearse, 1985, 73 (1−3), 273−284.
A.; Liu, B.; Wang, H.; Rubloff, G.; et al. Negating interfacial (25) Minami, T. Recent progress in superionic conducting glasses. J.
impedance in garnet-based solid-state Li metal batteries. Nat. Mater. Non-Cryst. Solids 1987, 95−96, 107−118.
2017, 16 (5), 572−579. (26) Minami, T.; Hayashi, A.; Tatsumisago, M. Recent progress of
(5) Janek, J.; Zeier, W. G. A solid future for battery development. glass and glass-ceramics as solid electrolytes for lithium secondary
Nat. Energy 2016, 1 (9), 16141. batteries. Solid State Ionics 2006, 177 (26−32), 2715−2720.
(6) Hu, Y.-S. Batteries: getting solid. Nat. Energy 2016, 1 (4), 16042. (27) Tatsumisago, M.; Hayashi, A. Superionic glasses and glass-
(7) Chen, R.; Li, Q.; Yu, X.; Chen, L.; Li, H. Approaching practically ceramics in the Li2S-P2S5 system for all-solid-state lithium secondary
accessible solid-state batteries: stability issues related to solid batteries. Solid State Ionics 2012, 225, 342−345.
electrolytes and interfaces. Chem. Rev. 2020, 120 (14), 6820−6877. (28) Yu, X.; Bates, J.; Jellison, G.; Hart, F. A stable thin-film lithium
(8) Banerjee, A.; Wang, X.; Fang, C.; Wu, E. A.; Meng, Y. S. electrolyte: lithium phosphorus oxynitride. J. Electrochem. Soc. 1997,
Interfaces and interphases in all-solid-state batteries with inorganic 144 (2), 524−532.
solid electrolytes. Chem. Rev. 2020, 120 (14), 6878−6933. (29) Mizuno, F.; Hayashi, A.; Tadanaga, K.; Tatsumisago, M. New,
(9) Manthiram, A.; Yu, X.; Wang, S. Lithium battery chemistries highly ion-conductive crystals precipitated from Li2S-P2S5 glasses.
enabled by solid-state electrolytes. Nat. Rev. Mater. 2017, 2 (4), Adv. Mater. 2005, 17 (7), 918−921.
16103. (30) Liu, Y.; Peng, H.; Su, H.; Zhong, Y.; Wang, X.; Xia, X.; Gu, C.;
(10) Zhao, Q.; Stalin, S.; Zhao, C.-Z.; Archer, L. A. Designing solid- Tu, J. Ultrafast Synthesis of I-Rich Lithium Argyrodite Glass-Ceramic
state electrolytes for safe, energy-dense batteries. Nat. Rev. Mater. Electrolyte with High Ionic Conductivity. Adv. Mater. 2022, 34 (3),
2020, 5 (3), 229−252. 2107346.
(11) Wang, S.; Zhang, W.; Chen, X.; Das, D.; Ruess, R.; Gautam, A.; (31) Famprikis, T.; Canepa, P.; Dawson, J. A.; Islam, M. S.;
Walther, F.; Ohno, S.; Koerver, R.; Zhang, Q.; et al. Influence of Masquelier, C. Fundamentals of inorganic solid-state electrolytes for
crystallinity of lithium thiophosphate solid electrolytes on the batteries. Nat. Mater. 2019, 18 (12), 1278−1291.
performance of solid-state batteries. Adv. Energy Mater. 2021, 11 (32) Wang, C.; Liang, J.; Kim, J. T.; Sun, X. Prospects of halide-
(24), 2100654. based all-solid-state batteries: From material design to practical
(12) Stramare, S.; Thangadurai, V.; Weppner, W. Lithium application. Sci. Adv. 2022, 8 (36), No. eadc9516.
lanthanum titanates: a review. Chem. Mater. 2003, 15 (21), 3974− (33) Zhou, L.; Zuo, T.-T.; Kwok, C. Y.; Kim, S. Y.; Assoud, A.;
3990. Zhang, Q.; Janek, J.; Nazar, L. F. High areal capacity, long cycle life 4
(13) Kamaya, N.; Homma, K.; Yamakawa, Y.; Hirayama, M.; Kanno, V ceramic all-solid-state Li-ion batteries enabled by chloride solid
R.; Yonemura, M.; Kamiyama, T.; Kato, Y.; Hama, S.; Kawamoto, K.; electrolytes. Nat. Energy 2022, 7 (1), 83−93.
et al. A lithium superionic conductor. Nat. Mater. 2011, 10 (9), 682− (34) Kwak, H.; Han, D.; Lyoo, J.; Park, J.; Jung, S. H.; Han, Y.;
686. Kwon, G.; Kim, H.; Hong, S. T.; Nam, K. W.; et al. New Cost-
(14) Kato, Y.; Hori, S.; Saito, T.; Suzuki, K.; Hirayama, M.; Mitsui, Effective Halide Solid Electrolytes for All-Solid-State Batteries:
A.; Yonemura, M.; Iba, H.; Kanno, R. High-power all-solid-state Mechanochemically Prepared Fe3+-Substituted Li2ZrCl6. Adv. Energy
batteries using sulfide superionic conductors. Nat. Energy 2016, 1 (4), Mater. 2021, 11 (12), 2003190.
16030. (35) Fu, J.; Yang, S.; Hou, J.; Azhari, L.; Yao, Z.; Ma, X.; Liu, Y.;
(15) Adeli, P.; Bazak, J. D.; Park, K. H.; Kochetkov, I.; Huq, A.; Vanaphuti, P.; Meng, Z.; Yang, Z.; et al. Modeling assisted synthesis of
Goward, G. R.; Nazar, L. F. Boosting solid-state diffusivity and Zr-doped Li3‑xIn1‑xZrxCl6 with ultrahigh ionic conductivity for lithium-
conductivity in lithium superionic argyrodites by halide substitution. ion batteries. J. Power Sources 2023, 556, 232465.
Angew. Chem., Int. Ed. 2019, 58 (26), 8681−8686. (36) Park, J.; Han, D.; Kwak, H.; Han, Y.; Choi, Y. J.; Nam, K.-W.;
(16) Asano, T.; Sakai, A.; Ouchi, S.; Sakaida, M.; Miyazaki, A.; Jung, Y. S. Heat treatment protocol for modulating ionic conductivity
Hasegawa, S. Solid halide electrolytes with high lithium-ion via structural evolution of Li3‑xYb1‑xMxCl6 (M = Hf4+, Zr4+) new
conductivity for application in 4 V class bulk-type all-solid-state halide superionic conductors for all-solid-state batteries. Chem. Eng. J.
batteries. Adv. Mater. 2018, 30 (44), 1803075. 2021, 425, 130630.
(17) Liang, J.; Li, X.; Wang, S.; Adair, K. R.; Li, W.; Zhao, Y.; Wang, (37) Kwak, H.; Han, D.; Son, J. P.; Kim, J. S.; Park, J.; Nam, K.-W.;
C.; Hu, Y.; Zhang, L.; Zhao, S.; et al. Site-occupation-tuned superionic Kim, H.; Jung, Y. S. Li+ conduction in aliovalent-substituted
LixScCl3+x halide solid electrolytes for all-solid-state batteries. J. Am. monoclinic Li2ZrCl6 for all-solid-state batteries: Li2+xZr1‑xMxCl6
Chem. Soc. 2020, 142 (15), 7012−7022. (M= In, Sc). Chem. Eng. J. 2022, 437, 135413.
(18) Li, X.; Liang, J.; Luo, J.; Norouzi Banis, M.; Wang, C.; Li, W.; (38) Chen, S.; Yu, C.; Chen, S.; Peng, L.; Liao, C.; Wei, C.; Wu, Z.;
Deng, S.; Yu, C.; Zhao, F.; Hu, Y.; et al. Air-stable Li3InCl6 electrolyte Cheng, S.; Xie, J. Enabling ultrafast lithium-ion conductivity of
with high voltage compatibility for all-solid-state batteries. Energy Li2ZrCl6 by indium doping. Chin. Chem. Lett. 2022, 33 (10), 4635−
Environ. Sci. 2019, 12 (9), 2665−2671. 4639.
(19) Auvergniot, J.; Cassel, A.; Ledeuil, J.-B.; Viallet, V.; Seznec, V.; (39) Hu, L.; Wang, J.; Wang, K.; Gu, Z.; Xi, Z.; Li, H.; Chen, F.;
Dedryvère, R. Interface stability of argyrodite Li6PS5Cl toward Wang, Y.; Li, Z.; Ma, C. A cost-effective, ionically conductive and
LiCoO2, LiNi1/3Co1/3Mn1/3O2, and LiMn2O4 in bulk all-solid-state compressible oxychloride solid-state electrolyte for stable all-solid-
batteries. Chem. Mater. 2017, 29 (9), 3883−3890. state lithium-based batteries. Nat. Commun. 2023, 14 (1), 3807.
(20) Yokokawa, H. Thermodynamic stability of sulfide electrolyte/ (40) Li, B.; Li, Y.; Zhang, H.-S.; Wu, T.-T.; Guo, S.; Cao, A.-M. Fast
oxide electrode interface in solid-state lithium batteries. Solid State Li+-conducting Zr4+-based oxychloride electrolyte with good thermal
Ionics 2016, 285, 126−135. and solvent stability. Sci. China Mater. 2023, 66, 3123−3128.
(21) Xu, S.; Hu, L. Towards a high-performance garnet-based solid- (41) Zhang, S.; Zhao, F.; Chen, J.; Fu, J.; Luo, J.; Alahakoon, S. H.;
state Li metal battery: A perspective on recent advances. J. Power Chang, L.-Y.; Feng, R.; Shakouri, M.; Liang, J.; et al. A family of
Sources 2020, 472, 228571. oxychloride amorphous solid electrolytes for long-cycling all-solid-
(22) Kato, Y.; Shiotani, S.; Morita, K.; Suzuki, K.; Hirayama, M.; state lithium batteries. Nat. Commun. 2023, 14 (1), 3780.
Kanno, R. All-solid-state batteries with thick electrode configurations. (42) Ishiguro, Y.; Ueno, K.; Nishimura, S.; Iida, G.; Igarashib, Y.
J. Phys. Chem. Lett. 2018, 9 (3), 607−613. TaCl5-glassified ultrafast lithium ion-conductive halide electrolytes for

27786 https://doi.org/10.1021/jacs.3c10602
J. Am. Chem. Soc. 2023, 145, 27774−27787
Journal of the American Chemical Society pubs.acs.org/JACS Article

high-performance all-solid-state lithium batteries. Chem. Lett. 2023, 52 Conductivity > 10 mS cm−1 for All-Solid-State Batteries. Angew.
(4), 237−241. Chem., Int. Ed. 2023, 62 (13), No. e202217581.
(43) Xu, R.; Yao, J.; Zhang, Z.; Li, L.; Wang, Z.; Song, D.; Yan, X.; (64) Stallworth, P.; Fontanella, J.; Wintersgill, M.; Scheidler, C. D.;
Yu, C.; Zhang, L. Room Temperature Halide-Eutectic Solid Immel, J. J.; Greenbaum, S.; Gozdz, A. NMR, DSC and high pressure
Electrolytes with Viscous Feature and Ultrahigh Ionic Conductivity. electrical conductivity studies of liquid and hybrid electrolytes. J.
Adv. Sci. 2022, 9 (35), 2204633. Power Sources 1999, 81−82, 739−747.
(44) Jung, S.-K.; Gwon, H.; Yoon, G.; Miara, L. J.; Lacivita, V.; Kim, (65) Huang, S.-D.; Shang, C.; Kang, P.-L.; Liu, Z.-P. Atomic
J.-S. Pliable lithium superionic conductor for all-solid-state batteries. structure of boron resolved using machine learning and global
ACS Energy Lett. 2021, 6 (5), 2006−2015. sampling. Chem. Sci. 2018, 9 (46), 8644−8655.
(45) Huang, S. D.; Shang, C.; Kang, P. L.; Zhang, X. J.; Liu, Z. P. (66) Steinhardt, P. J.; Nelson, D. R.; Ronchetti, M. Bond-
LASP: Fast global potential energy surface exploration. WIRES orientational order in liquids and glasses. Phys. Rev. B 1983, 28 (2),
Comput. Mol. Sci. 2019, 9 (6), No. e1415. 784−805.
(46) Shang, C.; Liu, Z.-P. Stochastic surface walking method for (67) Shang, C.; Zhang, X. J.; Liu, Z. P. Crystal phase transition of
structure prediction and pathway searching. J. Chem. Theory Comput. urea: what governs the reaction kinetics in molecular crystal phase
2013, 9 (3), 1838−1845. transitions. Phys. Chem. Chem. Phys. 2017, 19 (47), 32125−32131.
(47) Peng, Y.; Shang, C.; Liu, Z.-P. The dome of gold nanolized for (68) Huang, S.-D.; Shang, C.; Zhang, X.-J.; Liu, Z.-P. Material
catalysis. Chem. Sci. 2021, 12 (15), 5664−5671. discovery by combining stochastic surface walking global optimization
(48) Ma, S.; Huang, S.-D.; Liu, Z.-P. Dynamic coordination of with a neural network. Chem. Sci. 2017, 8 (9), 6327−6337.
cations and catalytic selectivity on zinc-chromium oxide alloys during (69) Maier, J. Nanoionics: ion transport and electrochemical storage
syngas conversion. Nat. Catal. 2019, 2 (8), 671−677. in confined systems. Nat. Mater. 2005, 4 (11), 805−815.
(49) Kang, P.-l.; Shang, C.; Liu, Z.-p. Recent implementations in (70) Zeng, D.; Yao, J.; Zhang, L.; Xu, R.; Wang, S.; Yan, X.; Yu, C.;
LASP 3.0: Global neural network potential with multiple elements Wang, L. Promoting favorable interfacial properties in lithium-based
and better long-range description. Chin. J. Chem. Phys. 2021, 34 (5), batteries using chlorine-rich sulfide inorganic solid-state electrolytes.
583−590. Nat. Commun. 2022, 13 (1), 1909.
(50) Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab (71) Arnold, W.; Buchberger, D. A.; Li, Y.; Sunkara, M.; Druffel, T.;
initio total-energy calculations using a plane-wave basis set. Phys. Rev. Wang, H. Halide doping effect on solvent-synthesized lithium
B 1996, 54 (16), 11169−11186. argyrodites Li6PS5X (X= Cl, Br, I) superionic conductors. J. Power
(51) Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized gradient Sources 2020, 464, 228158.
approximation made simple. Phys. Rev. Lett. 1996, 77 (18), 3865− (72) Sebti, E.; Evans, H. A.; Chen, H.; Richardson, P. M.; White, K.
3868. M.; Giovine, R.; Koirala, K. P.; Xu, Y.; Gonzalez-Correa, E.; Wang, C.;
(52) Ehrlich, S.; Moellmann, J.; Reckien, W.; Bredow, T.; Grimme, et al. Stacking Faults Assist Lithium-Ion Conduction in a Halide-
S. System-dependent dispersion coefficients for the DFT-D3 Based Superionic Conductor. J. Am. Chem. Soc. 2022, 144 (13),
treatment of adsorption processes on ionic surfaces. ChemPhysChem 5795−5811.
2011, 12 (17), 3414−3420. (73) Sun, H.-H.; Manthiram, A. Impact of microcrack generation
(53) Ravel, B.; Newville, M. ATHENA, ARTEMIS, HEPHAESTUS: and surface degradation on a nickel-rich layered Li
data analysis for X-ray absorption spectroscopy using IFEFFIT. J. [Ni0. 9Co0.05Mn0.05]O2 cathode for lithium-ion batteries. Chem.
Synchrotron Radiat. 2005, 12 (4), 537−541. Mater. 2017, 29 (19), 8486−8493.
(54) Rehr, J. J.; Albers, R. C. Theoretical approaches to x-ray
absorption fine structure. Rev. Mod. Phys. 2000, 72 (3), 621−654.
(55) Koningsberger, D. C.; Prins, R. X-ray Absorption: Principles,
Applications, Techniques of EXAFS, SEXAFS and XANES; John Wiley
and Sons Inc.: New York, NY, United States, 1987.
(56) Xu, J.; Xia, Y.; Li, Z.; Chen, H.; Wang, X.; Sun, Z.; Yin, W.
Multi-physics instrument: Total scattering neutron time-of-flight
diffractometer at China Spallation Neutron Source. Nucl. Instrum.
Methods Phys. Res. Sect. A 2021, 1013, 165642.
(57) Hart, P. W.; Sommerfeld, J. T. Cost estimation of specialty
chemicals from laboratory-scale prices. Cost Eng. 1997, 39 (3), 31.
(58) Li, F.; Cheng, X.; Lu, L.-L.; Yin, Y.-C.; Luo, J.-D.; Lu, G.; Meng,
Y.-F.; Mo, H.; Tian, T.; Yang, J.-T.; et al. Stable All-Solid-State
Lithium Metal Batteries Enabled by Machine Learning Simulation
Designed Halide Electrolytes. Nano Lett. 2022, 22 (6), 2461−2469.
(59) Yin, Y.-C.; Yang, J.-T.; Luo, J.-D.; Lu, G.-X.; Huang, Z.; Wang,
J.-P.; Li, P.; Li, F.; Wu, Y.-C.; Tian, T.; et al. A LaCl3-based lithium
superionic conductor compatible with lithium metal. Nature 2023,
616 (7955), 77−83.
(60) Park, K. H.; Oh, D. Y.; Choi, Y. E.; Nam, Y. J.; Han, L.; Kim, J.
Y.; Xin, H.; Lin, F.; Oh, S. M.; Jung, Y. S. Solution-processable glass
LiI-Li4SnS4 superionic conductors for all-solid-state Li-ion batteries.
Adv. Mater. 2016, 28 (9), 1874−1883.
(61) Zhao, F.; Alahakoon, S. H.; Adair, K.; Zhang, S.; Xia, W.; Li,
W.; Yu, C.; Feng, R.; Hu, Y.; Liang, J.; et al. An air-stable and Li-
metal-compatible glass-ceramic electrolyte enabling high-performance
all-solid-state Li metal batteries. Adv. Mater. 2021, 33 (8), 2006577.
(62) Croce, F.; Appetecchi, G.; Persi, L.; Scrosati, B. Nanocomposite
polymer electrolytes for lithium batteries. Nature 1998, 394 (6692),
456−458.
(63) Tanaka, Y.; Ueno, K.; Mizuno, K.; Takeuchi, K.; Asano, T.;
Sakai, A. New Oxyhalide Solid Electrolytes with High Lithium Ionic

27787 https://doi.org/10.1021/jacs.3c10602
J. Am. Chem. Soc. 2023, 145, 27774−27787

You might also like