Yinglei
Wu
*a,
Runze
Zhang
ab,
Qianjin
Huang
ab,
Jingjing
Wang
b,
Kaiyue
Jiang
bd,
Zhenying
Chen
bc,
Jinhui
Zhu
*b and
Xiaodong
Zhuang
*bd
aSchool of Chemistry and Chemical Engineering, Shanghai University of Engineering Science, Shanghai 201620, China. E-mail: wuyl@sues.edu.cn
bThe Soft2D Lab, State Key Laboratory of Metal Matrix Composites, Shanghai Key Laboratory of Electrical Insulation and Thermal Ageing, School of Chemistry and Chemical Engineering, Shanghai Jiao Tong University, Shanghai 200240, China. E-mail: zhujinhui1109@sjtu.edu.cn; zhuang@sjtu.edu.cn
cCollege of Chemistry, Zhengzhou University, Zhengzhou 450001, Henan, China
dFrontiers Science Center for Transformative Molecules, Zhang Jiang Institute for Advanced Study, Shanghai Jiao Tong University, Shanghai 201203, China
First published on 24th October 2024
A MoO2-doped Li5.5PS4.5Cl1.5 solid electrolyte with ionic conductivity of 12 mS cm−1 and an electrochemical window of 4.3 V vs. Li/Li+ was prepared, which enables a LiNi0.8Co0.1Mn0.1O2-based full cell to deliver a specific capacity of 194 mA h g−1 at 0.1C and retain 80% capacity after 3500 cycles at 1C.
Of the various SSEs, Li argyrodite of Li6PS5Cl is the most widely used due to its high σi (>1 mS cm−1), excellent ductility (low grain boundary resistance), and cost-effective raw materials that avoid rare elements.6 However, Li6PS5Cl suffers from poor electrochemical stability and requires enhanced σi.7 Substituting P, S, and Cl sites in Li6PS5Cl with other elements can modify the Li sublattice and increase S2−/Cl− site disorder, facilitating Li+ diffusion and improving σi.8,9 Additionally, doping elements such as In, Sn, Mg, F, and I can aid in the formation of a favorable solid electrolyte interphase (SEI), stabilizing the Li metal anode.7,10–12 Despite these advances, little research has been dedicated to improving the oxidation stability of Li6PS5Cl for compatibility with uncoated high-voltage cathodes.
In this study, we introduce dual-functional MoO2 into the crystal lattice of Cl-rich Li5.5PS4.5Cl1.5 SSE by substituting a portion of P2S5 raw material with MoO2 (1–3%). The optimized Li5.51P0.99Mo0.01S4.48O0.02Cl1.5 SSE exhibits higher σi (12.0 vs. 9.1 mS cm−1) and a wider electrochemical stability window (4.3 vs. 3.1 V vs. Li/Li+) compared to the Li5.5PS4.5Cl1.5 matrix. As a result, a full cell using this modified SSE and an uncoated single-crystal LiNi0.8Co0.1Mn0.1O2 (NCM811) cathode achieves a high specific capacity (194 mA h g−1 at 0.1C) and an ultralong cycling life (80% capacity retention after 3500 cycles at 1C). In situ Raman spectroscopy and ex situ X-ray photoelectron spectroscopy (XPS) confirm the excellent oxidation resistance of the MoO2-doped Li argyrodite.
A series of MoO2-doped Li5.5+xP1−xMoxS4.5−2xO2xCl1.5 (x = 0.01, 0.02, 0.03) and control Li5.5PS4.5Cl1.5 SSEs were synthesized through a solid-state process involving initial ball milling, followed by cold pressing, and finally annealing. The prepared Li5.5+xP1−xMoxS4.5−2xO2xCl1.5 SSEs exhibit an aggregation morphology with primary particle sizes ranging from a few microns to ten microns, with all elements distributed homogeneously. The MoO2 doping does not alter the morphology and elemental distribution of the SSEs (Fig. S1–S5, ESI†). The crystalline structures of the synthesized SSEs were analyzed using X-ray diffraction (XRD). As shown in Fig. 1a, all samples exhibit similar XRD patterns that align with Li7PS6 (PDF#34-0688), indicating a cubic F
3m space group. As the MoO2 content increases, impurity peaks from LiCl and Li3PO4 become more pronounced due to the limited solubility of Mo and O dopants. Additionally, the main peaks in the XRD patterns of Li5.5+xP1−xMoxS4.5−2xO2xCl1.5 shift slightly to smaller 2θ angles as the dopant concentration increases (Fig. 1b), indicating unit cell expansion.10 This expansion can be attributed to the substitution of smaller P5+ ions with larger Mo4+ ions (34 vs. 65 pm). The doping of lower-valence Mo4+ also increases the Li+ concentration within the lattice, improving σi.7
![]() | ||
| Fig. 1 (a) and (b) XRD patterns of Li5.5+xP1−xMoxS4.5−2xO2xCl1.5 (x = 0, 0.01, 0.02, 0.03). Rietveld XRD pattern refinement (c) and crystal structural diagram (d) of Li5.51P0.99Mo0.01S4.48O0.02Cl1.5. | ||
To further evaluate the structural changes in the SSE following MoO2 incorporation, Rietveld refinement of the XRD pattern of Li5.51P0.99Mo0.01S4.48O0.02Cl1.5 was performed (Fig. 1c and Table S1, ESI†). The calculated pattern matches the experimental data, confirming the successful doping of MoO2, forming a cubic argyrodite-type Li5.51P0.99Mo0.01S4.48O0.02Cl1.5 SSE with the F
3m space group. Fig. 1d illustrates the crystalline structure of the MoO2-doped SSE, where Mo atoms substitute P atoms at the 4b site, and O atoms replace S atoms at the 16e site.10 Furthermore, the 7Li magic angle spinning nuclear magnetic resonance spectra for both Li5.51P0.99Mo0.01S4.48O0.02Cl1.5 and Li5.5PS4.5Cl1.5 show a single peak without a shift (Fig. S6, ESI†), indicating similar Li chemical environments in both samples. This further confirms the successful substitution of P and S atoms by Mo and O atoms, consistent with the XRD results.
The chemical composition of the synthesized Li5.5+xP1−xMoxS4.5−2xO2xCl1.5 (x = 0, 0.01) SSEs was examined using XPS. The P 2p XPS spectra (Fig. S7a, ESI†) show characteristic peaks for PS43− at 131.8 and 132.7 eV in both samples.13 Additionally, a peak at 133.7 eV corresponds to PO43−, likely due to minor oxidation of PS43−.14 The S 2p (Fig. S7b, ESI†) and Mo 3d (Fig. S7c, ESI†) XPS spectra of the Li5.51P0.99Mo0.01S4.48O0.02Cl1.5 SSE confirm the presence of Mo–S bonds and the absence of Mo–O bonds,15 further verifying successful doping of Mo and O atoms into the lattice, replacing P and S atoms, respectively. Raman spectra of all prepared SSEs show characteristic peaks for PS43− at 198, 265, 428, 576, and 598 cm−1.16 Additionally, the Mo–S peak at 490 cm−1 is observed in the MoO2-doped SSEs, with its intensity increasing as the MoO2 content rises (Fig. S8, ESI†).15
The σi of the synthesized Li5.5+xP1−xMoxS4.5−2xO2xCl1.5 (x = 0, 0.01, 0.02, 0.03) SSEs was measured using electrochemical impedance spectroscopy (EIS). The resulting Nyquist plots and corresponding Arrhenius curves are displayed in Fig. S9 (ESI†) and Fig. 2a, respectively. Based on the calculations, the room temperature (RT, 25 °C) σi and activation energy (Ea) of all the SSEs were determined (Fig. 2b). The RT σi of the Li5.5PS4.5Cl1.5 matrix is calculated to be 9.1 mS cm−1, consistent with previously reported values.17 Notably, doping with 1% MoO2 significantly enhances the RT σi of the SSE to 12.0 mS cm−1. To the best of our knowledge, this is the highest reported RT σi, surpassing those of previously reported O-doped Li argyrodites (Fig. S10 and Table S2, ESI†), as O doping typically has a negative impact on the σi of SSEs.11 The enhanced RT σi of the Li5.51P0.99Mo0.01S4.48O0.02Cl1.5 SSE can be attributed to (i) the inherently high σi of the Li5.5PS4.5Cl1.5 matrix, and (ii) moderate Mo doping, which increases Li+ concentration and expands Li+ transport pathways.10 However, with higher MoO2 doping levels, the detrimental effects of O doping become predominant, resulting in a decrease in RT σi. The Ea of the SSEs exhibits an inverse relationship with their RT σi, with the Li5.51P0.99Mo0.01S4.48O0.02Cl1.5 SSE showing the lowest Ea of 0.28 eV, indicating the fastest Li+ transfer kinetics.
![]() | ||
| Fig. 2 Electrochemical properties of Li5.5+xP1−xMoxS4.5−2xO2xCl1.5 (x = 0, 0.01, 0.02, 0.03). (a) Arrhenius curves, (b) RT σi and Ea, (c) σe, and (d) ESWs. | ||
The electronic conductivity (σe) of SEs is typically used to assess their ability to suppress the growth of Li/Li-alloy dendrites.7 As shown in Fig. S11 (ESI†) and Fig. 2c, the σe of the Li5.51P0.99Mo0.01S4.48O0.02Cl1.5 SSE is calculated to be 1.27 × 10−8 S cm−1, which is significantly lower than that of the Li5.5PS4.5Cl1.5 matrix (4.33 × 10−8 S cm−1). This suggests that the MoO2-doped SSE forms a more stable interface with Li/Li-alloy anodes, which also can be clarified in the Li|Li (Fig. S12, ESI†) and Li|Cu cells (Fig. S13, ESI†). The electrochemical stability window (ESW), which reflects the oxidative stability of SEs, was measured using linear sweep voltammetry (LSV). As shown in Fig. 2d, the MoO2-doped SSEs exhibit a much wider ESW compared to Li5.5PS4.5Cl1.5 SSE (4.3 vs. 3.1 V vs. Li/Li+), indicating that the MoO2-doped SSEs are more stable when paired with high-voltage Li-layered oxide cathodes.
The performance of the synthesized SSEs was further evaluated in Li–In|NCM811 full cells, where the SSEs functioned as both the separator and a component of the composite cathode. All tests were conducted at RT under 35 MPa, with a cathode active material (CAM) loading of 5.6 mg cm−2. Fig. 3a presents the initial voltage profiles of the full cells using the four different SSEs at 0.1C (1C = 200 mA g−1). Notably, the cell using the Li5.51P0.99Mo0.01S4.48O0.02Cl1.5 SSE delivers the highest specific capacity of 194 mA h g−1 with a coulombic efficiency (CE) of 80.5%. In comparison, the cell with the Li5.5PS4.5Cl1.5 SSE achieves an initial specific capacity of 173 mA h g−1 and a CE of 75.5%. At a higher rate of 1C, the cell with 1% MoO2-doped SSE still exhibits the highest specific capacity of 134 mA h g−1 (Fig. 3b).
The long-term electrochemical performance of the cells with the four SSEs was further assessed through repeated galvanostatic charge–discharge cycling (Fig. 3c). All cells demonstrate high stability over 200 cycles. Among them, the cell with 1% MoO2-doped SSE maintains the highest specific capacity, followed by the cell with 2% MoO2-doped SSE. The undoped SSE-based cell comes next, while the cell with 3% MoO2-doped SSE shows the lowest specific capacity. This suggests a positive correlation between the electrochemical performance of the full cells and the RT σi of the adopted SSEs.
The rate capability of the full cells with the synthesized SSEs was also evaluated (Fig. 3d and Fig. S14, ESI†). The cell with Li5.51P0.99Mo0.01S4.48O0.02Cl1.5 SSE demonstrates the best rate performance, achieving specific capacities of 194, 181, 160, 134, 103, and 78 mA h g−1 at rates of 0.1C, 0.2C, 0.5C, 1C, 2C, and 3C, respectively. Additionally, the specific capacity recovered when the rate returns to 0.1C. Long-term cycling stability was also compared for the full cells using the four SSEs. As shown in Fig. 3e, the cell with Li5.51P0.99Mo0.01S4.48O0.02Cl1.5 SSE operates for over 5000 cycles at 1C and maintains 80% capacity retention after 3500 cycles. In contrast, the cells with 2% MoO2-doped, undoped, and 3% MoO2-doped SSEs achieve 80% capacity retention after 2100, 1600, and 600 cycles, respectively. Overall, the full cell with 1% MoO2-doped SSE exhibits significantly better electrochemical performance compared to previously reported O-doped Li argyrodite-based full cells (Table S3, ESI†).
To assess the interface evolution within the full cells, in situ EIS measurements were conducted on Li5.51P0.99Mo0.01S4.48O0.02Cl1.5 and Li5.5PS4.5Cl1.5 SSE-based full cells during the initial charge–discharge cycles. Nyquist plots (Fig. S15a, ESI†) and the corresponding equivalent circuit fitting values (Table S4, ESI†) reveal that the cathodic interfacial resistance (RCI) of the Li5.51P0.99Mo0.01S4.48O0.02Cl1.5 SSE-based cell increases from 20.8 to 108.7 Ω over the first cycle. In comparison, the Li5.5PS4.5Cl1.5 SSE-based cell exhibits a larger initial RCI of 70.4 Ω, which increases to 239.8 Ω after the first cycle (Fig. S16a, ESI†). These results demonstrate that the MoO2-doped SSE provides better stability with NCM811 than the undoped SSE. To gain further insight, the distribution of the relaxation time (DRT) diagrams was transformed from the Nyquist plots, providing detailed impedance information. As shown in Fig. S15b, c and S16b, c (ESI†), the DRT spectra exhibit five distinct peaks corresponding to different resistances: grain boundary (10−6 s), SEI/cathode electrolyte interphase (CEI) (10−5 s), charge transfer at the anode and cathode (10−4–10−3 and 10−2–10−1 s), and solid-state diffusion (100–101 s).18 The results confirm that the RCI, including CEI and cathodic charge transfer resistance, increases continuously during charging and discharging.
In situ Raman spectroscopy was employed to detect the compositional evolution of the cathodic interface during the first charge–discharge cycle. The Raman spectra collected at the Li5.51P0.99Mo0.01S4.48O0.02Cl1.5 SSE–NCM811 interface (Fig. 4a) show only the characteristic peaks of the PS43− tetrahedron and the Mo–S bond.16 Furthermore, the intensity of the PS43− main peak at 428 cm−1 remains constant during cycling (Fig. S17a, ESI†), indicating the exceptional stability of the MoO2-doped SSE under high voltage and oxidation conditions. In sharp contrast, the in situ Raman spectra for the Li5.5PS4.5Cl1.5 SSE–NCM811 interface (Fig. 4b) show a distinct S–S bond peak at 472 cm−1 in addition to the PS43− peaks.19 Moreover, the intensity of the PS43− peak decreases during cycling (Fig. S17b, ESI†), suggesting that the undoped SSE is prone to oxidation, forming high-valence S species.
To further evaluate the antioxidative durability of the synthesized SSEs, full cells cycled for 200 cycles were disassembled, and the composite cathodes were analyzed. The cycled cathode with Li5.51P0.99Mo0.01S4.48O0.02Cl1.5 SSE retains an intact and dense morphology, whereas the cathode with Li5.5PS4.5Cl1.5 SSE exhibits voids and cracks (Fig. S18, ESI†). The P 2p XPS spectrum of the cycled Li5.51P0.99Mo0.01S4.48O0.02Cl1.5 SSE (Fig. 4c and Table S5, ESI†) shows peaks corresponding to PS43− and PO43−,14 with the PO43− content remaining close to that in the fresh SSE. In contrast, the cycled Li5.5PS4.5Cl1.5 SSE exhibits an increased PO43− content and the formation of a new oxidized product, P2O5 (at 134.6 eV).14 The S 2p XPS spectrum of the cycled Li5.51P0.99Mo0.01S4.48O0.02Cl1.5 SSE (Fig. 4d and Table S5, ESI†) shows, in addition to the PS43− and S–Mo bond, a small amount of sulfate (5.44%).20 In contrast, around 40% of the Li5.5PS4.5Cl1.5 SSE is oxidized to form S–S bonds (163.5 and 164.7 eV) and sulfates after prolonged cycling.20 The Mo 3d XPS spectrum of the cycled Li5.51P0.99Mo0.01S4.48O0.02Cl1.5 SSE (Fig. 4e) further confirms the high stability of the Mo–S bond under high-voltage conditions. In conclusion, MoO2 doping significantly enhances the antioxidative stability of Li argyrodite, improving their performance in high-voltage applications.
In summary, a series of MoO2-doped Li argyrodites were successfully synthesized via the conventional solid-state reaction method. The optimized SSE, with 1% MoO2 doping, exhibits a high RT σi of 12.0 mS cm−1 and a broad ESW of 4.3 V vs. Li/Li+. These properties enable the corresponding Li–In|NCM811 full cell to achieve an impressive specific capacity of 194 mA h g−1 at 0.1C, along with an ultra-long cycling life, retaining 80% of its capacity after 3500 cycles at 1C. In situ EIS, in situ Raman, and ex situ XPS analyses further confirm that the MoO2-doped Li argyrodite demonstrates superior resistance to oxidation compared to the undoped counterpart.
This work was financially supported by the National Natural Science Foundation of China (NSFC: 52173205).
Footnote |
| † Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4cc04771k |
| This journal is © The Royal Society of Chemistry 2024 |