Next Article in Journal
Progress and Prospect of Practical Lithium-Sulfur Batteries Based on Solid-Phase Conversion
Next Article in Special Issue
Facile Synthesizing Yolk-Shelled Fe3O4@Carbon Nanocavities with Balanced Physiochemical Synergism as Efficient Hosts for High-Performance Lithium–Sulfur Batteries
Previous Article in Journal
Stabilizing a Zn Anode by an Ionic Amphiphilic Copolymer Electrolyte Additive for Long-Life Aqueous Zn-Ion Batteries
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Stabilizing of 1T-MoS2 for All-Solid-State Lithium-Ion Batteries

1
Fujian Key Laboratory of Electrochemical Energy Storage Materials, Fuzhou University, Fuzhou 350116, China
2
Agency for Science, Technology and Research (A*STAR), Innovis, Institute of Materials Research and Engineering, Singapore 138634, Singapore
*
Authors to whom correspondence should be addressed.
Batteries 2023, 9(1), 26; https://doi.org/10.3390/batteries9010026
Submission received: 21 November 2022 / Revised: 22 December 2022 / Accepted: 28 December 2022 / Published: 29 December 2022

Abstract

:
All-solid-state batteries (SSBs) are prospective candidates for a range of energy accumulation systems, delivering higher energy densities compared to batteries which use liquid electrolytes. Amongst the numerous solid-state electrolytes (SEs), sulfide-based electrolytes in particular have received more attention given that they have a high ionic conductivity. However, the incompatibility between the electrode and SEs is still an ongoing challenge that leads to poor electrochemical performance. In this work, we focus on 1T-MoS2. It is well known that 1T metallic MoS2 is unstable even at room temperature. However, we showed that 1T-MoS2 can be stabilized at 600 °C for at least 2 h, and the 1T-MoS2-600 interlayer spacing expanded to 0.95 nm. The high crystallinity of the 1T phase is highly compatible with solid electrolytes and coupled with the increased interlayer spacing, so in the all-solid-state lithium-ion battery (ALLLIB), we achieved outstanding cycling performance. At the current density of 0.2 C (1 C = 670 mA g−1), this material delivered a capacity of 406 mA h g−1 after 50 cycles.

1. Introduction

As renewable source of energy devices, lithium-ion batteries have been widely applied to replace traditional fossil resources [1,2,3,4,5,6,7]. Unfortunately, the combustibility of organic liquid electrolytes is a serious fire hazard which raises safety concerns [8,9,10,11,12,13]. To this end, solid electrolytes (SEs) were proposed as a solution due to their nonflammable characteristics and comparable ionic conductivity, especially in sulfide-based electrolytes, because they have a high ionic conductivity [14,15,16,17,18,19]. However, the compatibility between the electrode and SEs is still an ongoing challenge and impedes the progression of ASSLIB [20,21,22,23,24,25,26]. Selecting a suitable electrode material that is compatible with SEs has become essential and urgent for enhancing the electrochemical performance of ASSLIBs. Two-dimensional transition metal dichalcogenides (2D TMDs) such as 2D Molybdenum disulfide (MoS2) are regarded as promising candidates for ASSLIBs among the various electrode materials, owing to their intrinsically higher ionic conductivity, large layer spacing and compatibility with sulfide-based SEs [27,28,29]. The application of these materials in energy storage has been intensively researched such as sodium-ion batteries, ALLLIBs, zinc-ion batteries etc. [30,31,32,33]. However, their noticeable volume variation and the poor conductivity of electricity in charge/discharge process impedes further application of MoS2. Due to these constraints, nanostructural 1T-MoS2 which has a more stable electrochemical performance has been suggested as the solution. MoS2 has three different phases including 1T, 2H, and 3R. Many reports have researched 2H and 3R phases of MoS2. Additionally, a new problem arises, because it is well known that metastable phase 1T-MoS2 is unstable and degrades into its 2H even at room temperature. In previous works, Santhosha et al. [34] investigated exfoliated MoS2 as the electrode of ASSLIBs, the nanosized electrode materials could reduce the volume expansion and enhance the compatibility with SEs, so the capacity of 312 mA h g−1 at a current density of 0.1 C was maintained after 500 cycles. Li et al. [35] adopted O-doping and structure defects manufacturing to produce abundant 1T-MoS2(D-MoS2-O), metallic 1T-MoS2 can improve the conductivity and the expanded planes availed the diffusion of Zn2+, so they had tailored this material with long cycling durability and high-rate capability. Mirabal et al. [36] used a simple procedure to create PEO intercalation MoS2 composites, additionally, the increase electrochemical performance proved that interlayer expansion is a successful method for creating high-performance anode materials for ion store battery technologies.
In this work, we overcome the bottleneck of 1T-MoS2 degrading into its 2H phase even at high temperatures by using facile solvothermal synthesis method. By doing so, 1T-MoS2 with a stable phase can been obtained and the resulting compound remains phase-stable even at temperatures as high as 600 °C. Furthermore, the interlayer spacing of 1T-MoS2 expands to 0.95 nm, contrary to 1T-MoS2-200 is 0.62 nm. These expanded planes availed the diffusion of Li+ [37]. Furthermore, owing to that the higher crystallinity was favored for the compatibility between the electrode and SEs, our facile solvothermal synthesized 1T-MoS2 also exhibited a superior electrochemical performance. A capacity of 406 mA h g−1 has been achieved after 50 cycles at a current density of 0.2 C. Even at a larger current density of 0.5 C, our 1T-MoS2 electrode achieved a capacity of 250.2 mA h g−1, exhibiting superior performance.

2. Experimental Section

2.1. Synthesis of MoS2 Hollow Spheres

As with typical synthesis, 0.2 g of molybdenyl acetylacetonate was added into 30 mL of deionized water, stirring for about 30 min. Then, 0.4 g of thiourea was dissolved in the above solution. This reaction takes place at 200 °C for 12 h and is subsequently allowed to cool down naturally. After that, a black product was centrifuged and repeatedly rinsed with ethanol. Finally, the sample was dried at 70 °C under a vacuum overnight (1T-MoS2-200 at this stage). Following the above procedure, the samples were then heated at 600 °C for 2 h in an argon atmosphere resulting in 1T-MoS2-600.

2.2. Electrochemical Measurements

The working electrode was prepared by mixing the MoS2, acetylene black, and polyvinylidene fluoride (PVDF) (in a weight ratio of 7:2:1). Then, N-methylpyrrolidone (NMP) was used to dilute mixture to obtain a homogenous mud. ‘The mud’ was evenly spread on copper foil. Overnight, the foil was vacuum-dried at 110 °C. The sulfide-based solid electrolytes system was prepared by ball-milling 75% Li2S and 25% P2S5 (molar percent) in inert Argon gas for about 24 h. At the same time, the counter electrode and reference electrode were composed of LiIn. A typical cell device was prepared using a cylinder, and the cell was pressed at 35 tons/cm2 for 5 min after the enclosure was assembled in the glovebox. CV curves were captured in the working voltage range of 0.2–3 V at a scan rate of 0.2 mV s−1 (EC-lab, VMP-300), and charge–discharge profiles were collected on the electrochemical analyzer at 25 °C (Hokuto Denko, HJ-SM8, Tokyo, Japan).

2.3. Characterization of the Samples

X-ray diffraction (XRD) (Rigaku, RINT-2200, CuKα, Tokyo, Japan) data was collected in 5–80° to characterize the crystallinity of the samples. An HR Evolution Raman system was employed to obtain Raman spectra (532 nm laser). The morphology of samples was observed by scanning electron microscope (SEM, Hitachi S4800). X-ray photoelectron spectroscopy (XPS) was performed in the range of 0–1200 eV to investigate the states of Mo and S.

3. Results and Discussion

Figure 1a shows the XRD patterns of 1T-MoS2-200 and 1T-MoS2-600. There are two peaks in the XRD pattern at 9.1° and 18.4° of 1T-MoS2-200, which could be assigned to the (002) and (004) planes, respectively. Approximately 0.95 nm was determined as the interlayer spacing according to Bragg’s Law equation. From Figure 1a, we can see 1T-MoS2-600 has shifted XRD peaks (at 13.8° and 26.5°). This is due to high-temperature annealing step which increased the crystallinity, resulting in reduced interplanar spacing [38]. However, the existence of (004) reflection for 1T-MoS2-600 at 26.5° indicates that the high temperature did not degrade the 1T-MoS2 into its 2H-phase. To further support this claim, the Raman spectra of the two samples (Figure 1b) shows that the two samples share similar profiles. Specifically, a small peak at 375 cm−1 is observed for a mode of 2H-MoS2 in 1T-MoS2-600. A series of characteristic peaks at 110, 120, 144, 187, 207, 281, and 333 cm−1 can be assigned to 1T-MoS2 [39]. These results reveal that 1T phase 1T-MoS2-200 has not transitioned from to the 2H phase after annealing at 600 °C [40].
In order to compare the carbon content of 1T-MoS2-600 with 1T-MoS2-200, the thermogravimetric (TG) information of the composite material 1T-MoS2-600 and 1T-MoS2-200 are presented in Figure S1. From the TG curves, the drop in weight before 200 °C is attributed to the evaporation of residues of organic matter and moisture in material. In the range of 200 to 350 °C [41], oxidation of both 1T-MoS2-600 and 1T-MoS2-200 contributed to the continued decrease in weight. Additionally, from the TG curves, the peak at 350–420 °C is a result of the combustion of carbon. The carbon amount in final composites doubtless were around 2.68% (1T-MoS2-600) and 10.5% (1T-MoS2-200). (The final product is MoO3).
XPS spectra was also employed to confirm the 1T-MoS2 phase (Figure 2). The characteristic peaks of Mo and S was detected in the survey spectrum of 1T-MoS2-600 (Figure 2a). Figure 2b shows the distinct peaks of 1T-MoS2 at 232.21 eV and 228.21 eV, which could be attributed to Mo 3d3/2 and Mo 3d5/2, respectively. The binding energies at 233.01 and 229.33 eV result from the trace existence of 2H-MoS2. The binding energy at 226.42 eV is assigned to the feature peak of S 2p [42]. In comparison with the XPS spectra of 1T-MoS2-200 in Figure 2c, the results revealed that the transformation of the 1T phase is not significantly impacted by the high temperature, which is conform to the outcomes from Raman spectra and XRD.
SEM images were collected to penetrate the morphologies of MoS2 samples, as shown in Figure 3. From Figure 3a, we can clearly see the spherical morphology of 1T-MoS2-200. The enlarged SEM image in Figure 3b reveals that the microspheres are composed of nanosheets with porous structures, which could suppress the volume change problem in the lithiation/de-lithiation process [43]. Figure 3c,d show 1T-MoS2-600 after annealing. We can see that the 1T-MoS2 still maintains its spherical morphology even after annealing, showing that 1T-MoS2 can be stabilized at 600 °C.
Figure 4 displays TEM images of the 1T-MoS2-600. We can see that the 1T-MoS2-600 exhibits a micro-spherical morphology, composed of nanosheets. The presence of nanosheets is favorable for reducing the diffusion pathway of Li+ and can improve the electrochemical performance. In Figure 4b, 1T-MoS2-600 has a lattice fringe of 0.95 nm (according to XRD data), which should be assigned to the (002) plane. In the magnified TEM image, the thin 1T-MoS2-600 can be seen and it benefits the diffusion of Li+. Expanded interlayer spacing of 1T-MoS2-600 and the layered structure promotes the diffusion of lithium ion [44]. Our TEM EDX images (Figure 4c–f) also shows a homogeneous Mo, S, and C element distribution in 1T MoS2-600. Then, we tested 1T-MoS2-600 and 1T-MoS2-200 electrochemical performances. As shown in Figure 5a, a stable capacity of 406 mA h g−1 was obtained after 50 cycles for 1T-MoS2-600, while the capacity decayed quickly to 212.2 mA h g−1 after few cycles for 1T-MoS2-200 (Figure S2). Even at 0.1 C, the capacity of 1T-MoS2-200 quickly degraded to 190 mA h g−1 over a few cycles (Figure S3). These results imply that the increased crystallinity for 1T-MoS2-600 is leads to interfacial stability of ASSLIBs. The corresponding charge–discharge profiles of 1T-MoS2-600 at 0.2 C share a similar result, indicating a stable and reversible electrochemical process (Figure 5b). The rate performance of 1T-MoS2-600 has been checked at increasing current densities as shown in Figure 5c. At 0.1, 0.2, 0.5, and 1 C specific capacities of 538, 460, 376 and 296 mA h g−1 were retained, respectively. A capacity of 424 mA h g−1 is maintained as the current density was returned to 0.2 C, exhibiting a superior rate performance. Figure 5d is charge–discharge profiles at varied current densities. There are prominent plateaus at varied charge–discharge curves, indicating that 1T-MoS2-600 has a better compatibility with the SEs, owing to its stable electrochemical kinetics.
To have an insight into the electrochemical kinetics, at the sweep rate of 0.2 mV s−1, the curves were performed. As demonstrated in Figure 6a, a broad peak appeared at 0.9 V in the first cathodic process, this might be demonstrated by the development of a solid electrolyte interlayer (SEI) and the insertion of Li+ [45]. The cathodic peak at 0.2 V could be ascribed to the LixMoS2 being converted to Li2S and Mo [25]. The corresponding anodic peaks at 1.91 and 2.20 V are linked to the extraction of Li+ and manufacture of MoS2 [46]. In the subsequent cycles, the profiles are almost overlapping, signifying stable electrochemical kinetics. To investigate the capacitive contribution from the diffusion-controlled process and capacitance effect, CV curves at 0.2 to 1 mV s−1 of stepwise-increasing sweep rate for 1T-MoS2-600 is plotted (Figure 6b). The dominant effect can be illustrated based on the following expression (1):
i = a v b
in which i and v represented the peak current and sweep rates, respectively. a and b are constants. When the value of b approached 0.5, the diffusion-controlled mechanism mostly affected the electrochemical process. By contrast, it was controlled by a capacitance effect when b was close to 1. The computed b values for the reduction and oxidation peaks were 0.73 and 0.78, respectively, as shown in Figure 6c. These values indicate that the cathodic and anodic processes are separately dominated by the diffusion process and capacitance effect. The specific ratio from the capacitance and diffusion-controlled process was estimated according to the subsequent expression (2):
i = k 1 v + k 2 v 1 / 2  
where k1 and k2 represent the constant values, and i and v are the corresponding sweep rate and current. The capacitive contribution increased as the scan rates rose, as illustrated in Figure 6d. More specifically, the contribution percentage from capacitance effect reached 70% at 1 mV s−1. These findings demonstrate that, in comparison to a diffusion-controlled process, the pseudocapacitive effect has a greater impact on the electrochemical process.
Ex situ XRD was used to monitor the reaction procedure of 1T-MoS2-600 during the lithiation/de-lithiation process. Figure 7 illustrates the initial state of 1T-MoS2-600 with two typical peaks at 13.8° (002) and 32.6° (100). As the discharging process takes place, the (002) plane signal becomes weaker and disappears completely at 0.2 V [47]. The weakening of the (002) signal peak is the result of intercalation of Li+ and the conversion reaction of the 1T-MoS2-600. The peak did not appear in the succeeding charge process, suggesting that the 1T-MoS2-600 was already partially converted to the amorphous phase [48]. On the other hand, the peak of the (100) plane was recovered to a higher degree, corresponding to the de-lithiation and the reformation of MoS2 process. However, we can see from Figure 7 that the peak of (100) plane did not recover back to its initial position before discharge. This might be due to the fact that some lithium was left in the crystal structure of 1T-MoS2-600. The following equations serve as a summary of the reaction mechanism (3–4):
M o S 2 + x L i + + x e L i x M o S 2  
L i x M o S 2 + ( 4 x ) L i + + 4 e M o + 2 L i 2 S
To further understand the electrochemical process, the electrochemical impedance spectra of 1T-MoS2-600 were tested at various temperatures (313, 323, 333, and 343 K), as illustrated in Figure 8. Based on the Arrhenius equation:
k = A e E a / R T  
where k is the rate constant, R is the molar gas constant, T is the thermodynamic temperature, Ea is the apparent activation energy, and A is the pre-exponential factor.
We can linearize the equation to obtain the following expression:
I n 1 R c t = I n A 1 E a R T  
where Rct is the charge transfer resistance, A1 is a constant, R is the molar gas constant, T is the temperature, and Ea is the activation energy. We obtained values for Rct at different temperatures, and the Ea of 1T-MoS2-600 was determined to be 24.28 kJ mol−1, showing that the high crystalline 1T-phase of MoS2 favors the electrochemical kinetics in ASSLIBs [49,50], so this material can demonstrate an outstanding cycling stability.
We also used galvanostatic intermittent titration technique (GITT) to analyze the diffusion efficiency of lithium-ion during the electrochemical process. Figure 9 shows the outcomes of the GITT curves for the 1T-MoS2-600 and 1T-MoS2-200 at 0.1 C. The calculation of diffusion coefficient was based on the following formula (7):
D s = 4 π τ ( n m V m S ) 2 ( Δ E s Δ E t ) 2
where τ is the relaxation time, Δ E s and Δ E t come from steady voltage change and transient voltage variation in the discharge/charge process. We can see that 1T-MoS2-600 has a higher Li+ diffusion than 1T-MoS2-200, indicating that the high crystallinity and the expanded (002) planes availed the compatibility between the electrode and SEs and the diffusion of Li+ [51,52,53], so this material can demonstrate an outstanding cycling stability.
Figure 10, which displays the performance of a few sulfide-based electrodes, further demonstrates the advantages of 1T-MoS2-600 in the matter of capacity and cycling stability. An outstanding performance of 1T-MoS2-600 compared to similar electrode materials demonstrates its immense potential for SSBs applications [34,54,55].

4. Conclusions

The incompatibility between electrodes and SEs is still an ongoing challenge that leads to poor electrochemical performance. In this work, 1T-MoS2-600 with high crystallinity was employed as an electrode in ASSLIBs and increased interfacial stability with SEs in ASSLIBs. Additionally, we overcame the bottleneck of 1T-MoS2 degrading into its 2H phase at high temperatures by using a facile solvothermal synthesis method. We discovered and showed that this material demonstrated an outstanding cycling stability with a capacity of 406 mA h g−1 after 50 cycles at a current density of 0.2 C. Even at a larger current density of 0.5 C, our 1T-MoS2 electrode achieved a capacity of 250.2 mA h g−1, exhibiting superior performance. This work is also applicable to explore other electrode materials to solve interface issues in ASSLIBs.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/batteries9010026/s1, Figure S1: TG curves of 1T−MoS2−600 and 1T−MoS2−200; Figure S2: Cycling performance of 1T−MoS2−200 at 0.2C.; Figure S3: Cycling performance of 1T−MoS2−200 at 0.1 C.

Author Contributions

Conceptualization, Y.L., J.W. and M.W.; methodology, P.C.; software, J.W.; validation, Y.L., J.W. and M.W.; formal analysis, J.W.; Investigation, Z.Z. and K.W.; resources, J.W. and M.W.; data curation, P.C.; writing—original draft preparation, P.C.; writing—review and editing, W.Z., J.W. and M.W.; visualization, Y.L.; supervision, M.W.; project administration, J.W. and M.W.; funding acquisition, J.W. and M.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Project of Science and Technology of Fuzhou City (Grant No. 2021-Y-080). J.W. acknowledges the Agency for Science, Technology and Research (Central Research Fund Award).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors thanks for the financial support from the Project of Science and Technology of Fuzhou City (Grant No. 2021-Y-080). J.W. acknowledges the Agency for Science, Technology and Research (Central Research Fund Award).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Weiling, M.; Pfeiffer, F.; Baghernejad, M. Vibrational Spectroscopy Insight into the Electrode|electrolyte Interface/Interphase in Lithium Batteries. Adv. Energy Mater. 2022, 12, 2202504. [Google Scholar] [CrossRef]
  2. Xu, J.; Cai, X.; Cai, S.; Shao, Y.; Hu, C.; Lu, S.; Ding, S. High-Energy Lithium Ion Batteries: Recent Progress and A Promising Future in Applications. Energy Environ. Mater. 2022, 1–60. [Google Scholar] [CrossRef]
  3. Oh, H.J.; Kang, H.K.; Ahn, H.; Park, J.; Choi, J.; Kim, H.Y.; Lee, E.; Yeo, S.Y.; Choi, Y.O.; Yeang, B.J.; et al. Layered oxide cathode-inspired secondary hard carbon microsphere anode material for high-power and long-life rechargeable batteries. Chem. Eng. J. 2023, 454, 140252. [Google Scholar] [CrossRef]
  4. Wu, F.; Maier, J.; Yu, Y. Guidelines and trends for next-generation rechargeable lithium and lithium-ion batteries. Chem. Soc. Rev. 2020, 49, 1569–1614. [Google Scholar] [CrossRef]
  5. Min, X.; Xu, G.; Xie, B.; Guan, P.; Sun, M.; Cui, G. Challenges of prelithiation strategies for next generation high energy lithium-ion batteries. Energy Stor. Mater. 2022, 47, 297–318. [Google Scholar] [CrossRef]
  6. Gao, H.; Neale, A.R.; Zhu, Q.; Bahri, M.; Wang, X.; Yang, H.; Xu, Y.; Clowes, R.; Browning, N.D.; Little, M.A.; et al. A Pyrene-4,5,9,10-Tetraone-Based Covalent Organic Framework Delivers High Specific Capacity as a Li-Ion Positive Electrode. J. Am. Chem. Soc. 2022, 144, 9434–9442. [Google Scholar] [CrossRef]
  7. Gao, H.; Tian, B.; Yang, H.; Neale, A.R.; Little, M.A.; Sprick, R.S.; Hardwick, L.J.; Cooper, A.I. Crosslinked Polyimide and Reduced Graphene Oxide Composites as Long Cycle Life Positive Electrode for Lithium-Ion Cells. ChemSusChem 2020, 13, 5571–5579. [Google Scholar] [CrossRef]
  8. Lewis, J.A.; Cavallaro, K.A.; Liu, Y.; McDowell, M.T. The promise of alloy anodes for solid-state batteries. Joule 2022, 6, 1418–1430. [Google Scholar] [CrossRef]
  9. Wang, L.; Sun, X.; Ma, J.; Chen, B.; Li, C.; Li, J.; Chang, L.; Yu, X.; Chan, T.-S.; Hu, Z.; et al. Bidirectionally Compatible Buffering Layer Enables Highly Stable and Conductive Interface for 4.5 V Sulfide-Based All-Solid-State Lithium Batteries. Adv. Energy Mater. 2021, 11, 2100881. [Google Scholar] [CrossRef]
  10. Cao, D.; Zhao, Y.; Sun, X.; Natan, A.; Wang, Y.; Xiang, P.; Wang, W.; Zhu, H. Processing Strategies to Improve Cell-Level Energy Density of Metal Sulfide Electrolyte-Based All-Solid-State Li Metal Batteries and Beyond. ACS Energy Lett. 2020, 5, 3468–3489. [Google Scholar] [CrossRef]
  11. Wang, J.; Okabe, J.; Urita, K.; Moriguchi, I.; Wei, M. Cu2S hollow spheres as an anode for high-rate sodium storage performance. J. Electroanal. Chem. 2020, 874, 114523. [Google Scholar] [CrossRef]
  12. Oh, P.; Yun, J.; Choi, J.H.; Saqib, K.S.; Embleton, T.J.; Park, S.; Lee, C.; Ali, J.; Ko, K.; Cho, J. Development of High-Energy Anodes for All-Solid-State Lithium Batteries Based on Sulfide Electrolytes. Angew. Chem. Int. Ed. 2022, 61, e202201249. [Google Scholar] [CrossRef] [PubMed]
  13. Wang, L.; Xie, R.; Chen, B.; Yu, X.; Ma, J.; Li, C.; Hu, Z.; Sun, X.; Xu, C.; Dong, S.; et al. In-situ visualization of the space-charge-layer effect on interfacial lithium-ion transport in all-solid-state batteries. Nat. Commun. 2020, 11, 5889. [Google Scholar] [CrossRef] [PubMed]
  14. Wu, Y.; Zhou, K.; Ren, F.; Ha, Y.; Liang, Z.; Zheng, X.; Wang, Z.; Yang, W.; Zhang, M.; Luo, M.; et al. Highly reversible Li2RuO3 cathodes in sulfide-based all solid-state lithium batteries. Energy Environ. Sci. 2022, 15, 3470–3482. [Google Scholar] [CrossRef]
  15. Fan, Z.; Xiang, J.; Yu, Q.; Wu, X.; Li, M.; Wang, X.; Xia, X.; Tu, J. High Performance Single-Crystal Ni-Rich Cathode Modification via Crystalline LLTO Nanocoating for All-Solid-State Lithium Batteries. ACS Appl. Mater. Interfaces 2022, 14, 726–735. [Google Scholar] [CrossRef] [PubMed]
  16. Yang, S.; Yamamoto, K.; Mei, X.; Sakuda, A.; Uchiyama, T.; Watanabe, T.; Takami, T.; Hayashi, A.; Tatsumisago, M.; Uchimoto, Y. High Rate Capability from a Graphite Anode through Surface Modification with Lithium Iodide for All-Solid-State Batteries. ACS Appl. Energy Mater. 2022, 5, 667–673. [Google Scholar] [CrossRef]
  17. Tan, D.H.S.; Meng, Y.S.; Jang, J. Scaling up high-energy-density sulfidic solid-state batteries: A lab-to-pilot perspective. Joule 2022, 6, 1755–1769. [Google Scholar] [CrossRef]
  18. Ohashi, A.; Kodama, M.; Xueying, S.; Hori, S.; Suzuki, K.; Kanno, R.; Hirai, S. Stress distribution in the composite electrodes of sulfide all-solid-state lithium-ion batteries. J. Power Sources 2020, 470, 228437. [Google Scholar] [CrossRef]
  19. Gregory, G.L.; Gao, H.; Liu, B.; Gao, X.; Rees, G.J.; Pasta, M.; Bruce, P.G.; Williams, C.K. Buffering Volume Change in Solid-State Battery Composite Cathodes with CO2-Derived Block Polycarbonate Ethers. J. Am. Chem. Soc. 2022, 144, 17477–17486. [Google Scholar] [CrossRef] [PubMed]
  20. Li, L.; Duan, H.; Zhang, L.; Deng, Y.; Chen, G. Optimized functional additive enabled stable cathode and anode interfaces for high-voltage all-solid-state lithium batteries with significantly improved cycling performance. J. Mater. Chem. A 2022, 10, 20331–20342. [Google Scholar] [CrossRef]
  21. Ye, L.; Li, X. A dynamic stability design strategy for lithium metal solid state batteries. Nature 2021, 593, 218–222. [Google Scholar] [CrossRef] [PubMed]
  22. Sun, Z.; Lai, Y.; Lv, N.; Hu, Y.; Li, B.; Jing, S.; Jiang, L.; Jia, M.; Li, J.; Chen, S.; et al. Boosting the Electrochemical Performance of All-Solid-State Batteries with Sulfide Li6PS5Cl Solid Electrolyte Using Li2WO4-Coated LiCoO2 Cathode. Adv. Mater. Interfaces 2021, 8, 2100624. [Google Scholar] [CrossRef]
  23. Wan, H.; Zhang, J.; Xia, J.; Ji, X.; He, X.; Liu, S.; Wang, C. F and N Rich Solid Electrolyte for Stable All-Solid-State Battery. Adv. Funct. Mater. 2022, 32, 2110876. [Google Scholar] [CrossRef]
  24. Li, Q.; Cao, Y.; Yin, G.; Gao, Y. The stable cycling of a high-capacity Bi anode enabled by an in situ-generated Li3PO4 transition layer in a sulfide-based all-solid-state battery. ChemComm 2020, 56, 15458–15461. [Google Scholar] [CrossRef] [PubMed]
  25. Priyadharsini, N.; Rupa Kasturi, P.; Shanmugavani, A.; Surendran, S.; Shanmugapriya, S.; Kalai Selvan, R. Effect of chelating agent on the sol-gel thermolysis synthesis of LiNiPO4 and its electrochemical properties for hybrid capacitors. J. Phys. Chem. Solids 2018, 119, 183–192. [Google Scholar] [CrossRef]
  26. Priyadharsini, N.; Shanmugapriya, S.; Kasturi, P.R.; Surendran, S.; Selvan, R.K. Morphology-dependent electrochemical properties of sol-gel synthesized LiCoPO4 for aqueous hybrid capacitors. Electrochim. Acta 2018, 289, 516–526. [Google Scholar] [CrossRef]
  27. Gao, Y.; Hai, P.; Liu, L.; Yin, J.; Gan, Z.; Ai, W.; Wu, C.; Cheng, Y.; Xu, X. Balanced Crystallinity and Nanostructure for SnS2 Nanosheets through Optimized Calcination Temperature toward Enhanced Pseudocapacitive Na+ Storage. ACS Nano 2022, 16, 14745–14753. [Google Scholar] [CrossRef]
  28. Geng, P.; Zheng, S.; Tang, H.; Zhu, R.; Zhang, L.; Cao, S.; Xue, H.; Pang, H. Transition Metal Sulfides Based on Graphene for Electrochemical Energy Storage. Adv. Energy Mater. 2018, 8, 1703259. [Google Scholar] [CrossRef]
  29. Priyadharsini, N.; Shanmugavani, A.; Surendran, S.; Senthilkumar, B.; Vasylechko, L.; Kalai Selvan, R. Improved electrochemical performances of LiMnPO4 synthesized by a hydrothermal method for Li-ion supercapatteries. J. Mater. Sci. Mater. Electron. 2018, 29, 18553–18565. [Google Scholar] [CrossRef]
  30. Zhang, Y.; Zhang, L.; Lv, T.A.; Chu, P.K.; Huo, K. Two-Dimensional Transition Metal Chalcogenides for Alkali Metal Ions Storage. ChemSusChem 2020, 13, 1114–1154. [Google Scholar] [CrossRef]
  31. Yin, X.; Tang, C.S.; Zheng, Y.; Gao, J.; Wu, J.; Zhang, H.; Chhowalla, M.; Chen, W.; Wee, A.T.S. Recent developments in 2D transition metal dichalcogenides: Phase transition and applications of the (quasi-)metallic phases. Chem. Soc. Rev. 2021, 50, 10087–10115. [Google Scholar] [CrossRef] [PubMed]
  32. Sun, Q.; Li, D.; Dai, L.; Liang, Z.; Ci, L. Structural Engineering of SnS2 Encapsulated in Carbon Nanoboxes for High-Performance Sodium/Potassium-Ion Batteries Anodes. Small 2020, 16, 2005023. [Google Scholar] [CrossRef] [PubMed]
  33. Wang, H.; Yuan, Q.; Wang, D.; Chen, G.; Cheng, X.; Kups, T.; Schaaf, P. Disordered surface formation of WS2via hydrogen plasma with enhanced anode performances for lithium and sodium ion batteries. Sustain. Energy Fuels 2019, 3, 865–874. [Google Scholar] [CrossRef]
  34. Santhosha, A.L.; Nayak, P.K.; Pollok, K.; Langenhorst, F.; Adelhelm, P. Exfoliated MoS2 as Electrode for All-Solid-State Rechargeable Lithium-Ion Batteries. J. Phys. Chem. C 2019, 123, 12126–12134. [Google Scholar] [CrossRef]
  35. Li, S.; Liu, Y.; Zhao, X.; Cui, K.; Shen, Q.; Li, P.; Qu, X.; Jiao, L. Molecular Engineering on MoS2 Enables Large Interlayers and Unlocked Basal Planes for High-Performance Aqueous Zn-Ion Storage. Angew. Chem. Int. Ed. 2021, 60, 20286–20293. [Google Scholar] [CrossRef]
  36. Mirabal, N.; Aguirre, P.; Santa Ana, M.A.; Benavente, E.; González, G. Thermal stability and electrical conductivity in polyethers-molybdenum disulfide nanocomposites. Electrochim. Acta 2003, 48, 2123–2127. [Google Scholar] [CrossRef]
  37. Wang, Y.; Yang, Y.; Zhang, D.; Wang, Y.; Luo, X.; Liu, X.; Kim, J.-K.; Luo, Y. Inter-overlapped MoS2/C composites with large-interlayer-spacing for high-performance sodium-ion batteries. Nanoscale Horiz. 2020, 5, 1127–1135. [Google Scholar] [CrossRef]
  38. Bai, J.; Zhao, B.; Zhou, J.; Si, J.; Fang, Z.; Li, K.; Ma, H.; Dai, J.; Zhu, X.; Sun, Y. Glucose-Induced Synthesis of 1T-MoS2/C Hybrid for High-Rate Lithium-Ion Batteries. Small 2019, 15, 1805420. [Google Scholar] [CrossRef]
  39. Zhuo, Y.; Prestat, E.; Kinloch, I.A.; Bissett, M.A. Self-Assembled 1T-MoS2/Functionalized Graphene Composite Electrodes for Supercapacitor Devices. ACS Appl. Energy Mater 2022, 5, 61–70. [Google Scholar] [CrossRef]
  40. Jiang, Y.; Li, S.; Zhang, F.; Zheng, W.; Zhao, L.; Feng, Q. Metal-semiconductor 1T/2H-MoS2 by a heteroatom-doping strategy for enhanced electrocatalytic hydrogen evolution. Catal. Commun. 2021, 156, 106325. [Google Scholar] [CrossRef]
  41. Fayed, M.G.; Attia, S.Y.; Barakat, Y.F.; El-Shereafy, E.E.; Rashad, M.M.; Mohamed, S.G. Carbon and nitrogen co-doped MoS2 nanoflakes as an electrode material for lithium-ion batteries and supercapacitors. SM&T 2021, 29, e00306. [Google Scholar]
  42. Acerce, M.; Voiry, D.; Chhowalla, M. Metallic 1T phase MoS2 nanosheets as supercapacitor electrode materials. Nat. Nanotechnol. 2015, 10, 313–318. [Google Scholar] [CrossRef] [PubMed]
  43. Liu, B.; Wang, L.; Zhu, Y.; Peng, H.; Du, C.; Yang, X.; Zhao, Q.; Hou, J.; Cao, C. Ammonium-Modified Synthesis of Vanadium Sulfide Nanosheet Assemblies toward High Sodium Storage. ACS Nano 2022, 16, 12900–12909. [Google Scholar] [CrossRef] [PubMed]
  44. Zhang, X.; Weng, W.; Gu, H.; Hong, Z.; Xiao, W.; Wang, F.; Li, W.; Gu, D. Versatile Preparation of Mesoporous Single-Layered Transition-Metal Sulfide/Carbon Composites for Enhanced Sodium Storage. Adv. Mater. 2022, 34, 2104427. [Google Scholar] [CrossRef] [PubMed]
  45. Yu, X.; Zhao, G.; Huang, H.; Liu, C.; Lyu, P.; Zhang, N. Interlayer-expanded MoS2 nanoflowers anchored on the graphene: A high-performance Li+/Mg2+ co-intercalation cathode material. Chem. Eng. J. 2022, 428, 131214. [Google Scholar] [CrossRef]
  46. Lu, H.; Tian, K.; Bu, L.; Huang, X.; Li, X.; Zhao, Y.; Wang, F.; Bai, J.; Gao, L.; Zhao, J. Synergistic effect from coaxially integrated CNTs@MoS2/MoO2 composite enables fast and stable lithium storage. J. Energy Chem. 2021, 55, 449–458. [Google Scholar] [CrossRef]
  47. Li, S.; Liu, Y.; Zhao, X.; Shen, Q.; Zhao, W.; Tan, Q.; Zhang, N.; Li, P.; Jiao, L.; Qu, X. Sandwich-Like Heterostructures of MoS2/Graphene with Enlarged Interlayer Spacing and Enhanced Hydrophilicity as High-Performance Cathodes for Aqueous Zinc-Ion Batteries. Adv. Mater. 2021, 33, 2007480. [Google Scholar] [CrossRef]
  48. Gong, F.; Ye, S.; Liu, M.; Zhang, J.; Gong, L.; Zeng, G.; Meng, E.; Su, P.; Xie, K.; Zhang, Y.; et al. Boosting electrochemical oxygen evolution over yolk-shell structured O–MoS2 nanoreactors with sulfur vacancy and decorated Pt nanoparticles. Nano Energy 2020, 78, 105284. [Google Scholar] [CrossRef]
  49. Zhou, L.-F.; Gao, X.-W.; Du, T.; Gong, H.; Liu, L.-Y.; Luo, W.-B. New Phosphate Zn2Fe(PO4)2 Cathode Material for Nonaqueous Zinc Ion Batteries with Long Life Span. ACS Appl. Mater. Interfaces 2022, 14, 8888–8895. [Google Scholar] [CrossRef] [PubMed]
  50. Wang, J.; Huang, J.; Huang, S.; Komine, Y.; Notohara, H.; Urita, K.; Moriguchi, I.; Wei, M. Regulating the effects of SnS shrinkage in all-solid-state lithium-ion batteries with excellent electrochemical performance. Chem. Eng. J. 2022, 429, 132424. [Google Scholar] [CrossRef]
  51. Wu, D.; Sun, F.; Qu, Z.; Wang, H.; Lou, Z.; Wu, B.; Zhao, G. Multi-scale structure optimization of boron-doped hard carbon nanospheres boosting the plateau capacity for high performance sodium ion batteries. J. Mater. Chem. A 2022, 10, 17225–17236. [Google Scholar] [CrossRef]
  52. Zhang, W.; Li, H.; Zhang, Z.; Xu, M.; Lai, Y.; Chou, S.-L. Full Activation of Mn4+/Mn3+ Redox in Na4MnCr(PO4)3 as a High-Voltage and High-Rate Cathode Material for Sodium-Ion Batteries. Small 2020, 16, 2001524. [Google Scholar] [CrossRef] [PubMed]
  53. Tao, L.; Yang, Y.; Wang, H.; Zheng, Y.; Hao, H.; Song, W.; Shi, J.; Huang, M.; Mitlin, D. Sulfur-nitrogen rich carbon as stable high capacity potassium ion battery anode: Performance and storage mechanisms. Energy Stor. Mater. 2020, 27, 212–225. [Google Scholar] [CrossRef]
  54. Oh, D.Y.; Choi, Y.E.; Kim, D.H.; Lee, Y.-G.; Kim, B.-S.; Park, J.; Sohn, H.; Jung, Y.S. All-solid-state lithium-ion batteries with TiS2 nanosheets and sulphide solid electrolytes. J. Mater. Chem. A 2016, 4, 10329–10335. [Google Scholar] [CrossRef]
  55. Wang, J.; Okabe, J.; Komine, Y.; Notohara, H.; Urita, K.; Moriguchi, I.; Wei, M. The optimized interface engineering of VS2 as cathodes for high performance all-solid-state lithium-ion battery. Sci. China Technol. Sci. 2022, 65, 1859–1866. [Google Scholar] [CrossRef]
Figure 1. (a) XRD patterns, and (b) Raman spectra of 1T−MoS2−200 and 1T−MoS2−600, respectively.
Figure 1. (a) XRD patterns, and (b) Raman spectra of 1T−MoS2−200 and 1T−MoS2−600, respectively.
Batteries 09 00026 g001
Figure 2. (a) XPS survey spectrum of 1T-MoS2-600, (b) Mo 3d spectrum of 1T-MoS2-600, (c) Mo 3d of 1T-MoS2-200, and (d) S 2p spectrum of 1T-MoS2-600.
Figure 2. (a) XPS survey spectrum of 1T-MoS2-600, (b) Mo 3d spectrum of 1T-MoS2-600, (c) Mo 3d of 1T-MoS2-200, and (d) S 2p spectrum of 1T-MoS2-600.
Batteries 09 00026 g002
Figure 3. SEM images of (a,b)1T-MoS2-200 and (c,d) 1T-MoS2-600.
Figure 3. SEM images of (a,b)1T-MoS2-200 and (c,d) 1T-MoS2-600.
Batteries 09 00026 g003
Figure 4. (a,b) TEM images and (cf) STEM image and elemental mapping images of 1T-MoS2-600.
Figure 4. (a,b) TEM images and (cf) STEM image and elemental mapping images of 1T-MoS2-600.
Batteries 09 00026 g004
Figure 5. (a) Cycling performance of 1T−MoS2−600 at 0.2 C, (b) the corresponding charge–discharge profiles at 0.2 C, (c) rate performance of 1T−MoS2−600, and (d) the related charge–discharge profiles at varied current densities from 0.1 to 1 C.
Figure 5. (a) Cycling performance of 1T−MoS2−600 at 0.2 C, (b) the corresponding charge–discharge profiles at 0.2 C, (c) rate performance of 1T−MoS2−600, and (d) the related charge–discharge profiles at varied current densities from 0.1 to 1 C.
Batteries 09 00026 g005
Figure 6. (a) CV curves of 1T−MoS2−600 at a sweep rate of 0.5 mV s−1, (b) CV curves of 1T−MoS2−600 at different sweep rates, (c) b values from the oxidation and reduction peaks, and (d) the pseudocapacitive contribution of 1T−MoS2−600.
Figure 6. (a) CV curves of 1T−MoS2−600 at a sweep rate of 0.5 mV s−1, (b) CV curves of 1T−MoS2−600 at different sweep rates, (c) b values from the oxidation and reduction peaks, and (d) the pseudocapacitive contribution of 1T−MoS2−600.
Batteries 09 00026 g006
Figure 7. Ex situ XRD patterns of 1T-MoS2-600.
Figure 7. Ex situ XRD patterns of 1T-MoS2-600.
Batteries 09 00026 g007
Figure 8. Electrochemical impedance spectra of 1T-MoS2-600 at an increasing temperature of 313, 323, 333, and 343 K, respectively.
Figure 8. Electrochemical impedance spectra of 1T-MoS2-600 at an increasing temperature of 313, 323, 333, and 343 K, respectively.
Batteries 09 00026 g008
Figure 9. (a) The discharge–charge GITT plots at 0.1 C, and (b) lithium−ion diffusion coefficient of 1T−MoS2−600 and 1T−MoS2−200 at the discharge and charge processes, respectively.
Figure 9. (a) The discharge–charge GITT plots at 0.1 C, and (b) lithium−ion diffusion coefficient of 1T−MoS2−600 and 1T−MoS2−200 at the discharge and charge processes, respectively.
Batteries 09 00026 g009
Figure 10. The cycling performance comparison of some sulfide-based electrode in ALLLIBs.
Figure 10. The cycling performance comparison of some sulfide-based electrode in ALLLIBs.
Batteries 09 00026 g010
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Chong, P.; Zhou, Z.; Wang, K.; Zhai, W.; Li, Y.; Wang, J.; Wei, M. The Stabilizing of 1T-MoS2 for All-Solid-State Lithium-Ion Batteries. Batteries 2023, 9, 26. https://doi.org/10.3390/batteries9010026

AMA Style

Chong P, Zhou Z, Wang K, Zhai W, Li Y, Wang J, Wei M. The Stabilizing of 1T-MoS2 for All-Solid-State Lithium-Ion Batteries. Batteries. 2023; 9(1):26. https://doi.org/10.3390/batteries9010026

Chicago/Turabian Style

Chong, Peidian, Ziwang Zhou, Kaihong Wang, Wenhao Zhai, Yafeng Li, Jianbiao Wang, and Mingdeng Wei. 2023. "The Stabilizing of 1T-MoS2 for All-Solid-State Lithium-Ion Batteries" Batteries 9, no. 1: 26. https://doi.org/10.3390/batteries9010026

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop