Next Article in Journal
Smart Design of Cz-Ge Crystal Growth Furnace and Process
Previous Article in Journal
Recent Progress in Phase Stability and Elastic Anomalies of Group VB Transition Metals
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Facile One-Pot Hydrothermal Synthesis of Hierarchical MoS2/α-MnS Nanocomposites with Good Cycling Performance as Anode Materials for Lithium-Ion Batteries

College of Sciences, Northeastern University, Shenyang 110819, China
*
Author to whom correspondence should be addressed.
Crystals 2022, 12(12), 1763; https://doi.org/10.3390/cryst12121763
Submission received: 1 November 2022 / Revised: 1 December 2022 / Accepted: 1 December 2022 / Published: 5 December 2022

Abstract

:
The design of nanoscale composites with a hierarchical structure can improve the poor cycling performances of transition metal sulfides such as anode materials. The hierarchical MoS2/α-MnS nanocomposites were synthesized by a facile one-pot hydrothermal method in this study, aiming to improve the cycling performance. As an anode material, hierarchical MoS2/α-MnS nanocomposites deliver a high reversible capacity, of 1498 mAh g−1 at a current density of 50 mA g−1; a capacity of ~800 mAh g−1 is maintained over 100 cycles at 300 mA g−1, and a capacity of 907 mAh g−1 is obtained at 1200 mA g−1. Moreover, these capacities increase with the number of cycles, which is mainly owed to the occurring metallic nanoparticles that catalyze the developing polymeric film and enhance the conductivity of the active materials during the electrochemical reactions. The good cycling performances are attributed to the synergistic effects between MoS2 and α-MnS.

1. Introduction

Electric vehicles (EVs) have become increasingly popular due to their environmental advantages. However, the relatively low energy density of the lithium-ion batteries (LIBs) used as power sources for the EVs in today’s market leads to range anxiety, to some extent, which results in many people having difficulty adopting EVs [1,2,3]. Developing high-performance LIBs with high energy density, low cost, long cycle life, and environmental friendliness is urgent. Transition metal sulfides are promising electrode materials because of their high theoretical lithium storage capacity, low cost, and nontoxicity to the environment [4,5,6,7]. The reaction mechanism between transition metal sulfides and lithium generally consists of two processes: insertion followed by conversion reaction. The insertion mechanism is the reaction of nLi + + ne + M x S y Li n M x S y ; the conversion mechanism can be described as Li n M x S y + ( 2 y n ) ( Li + + e - ) xM + yLi 2 S and Li 2 S S + 2 ( Li + + e ) [8,9,10,11,12,13,14,15,16,17,18,19,20,21]. The transition metal sulfides used as the anode materials suffer from the poor cycling performances due to the high solubility of the intermediates (lithium polysulfides) and the huge volume change during the charging/discharging process. The strategies, compositing and nanometerization, are widely applied to address these issues. The composites, which are comprised of conductive carbon, polymer, and transition metal sulfides, have been synthesized and deliver good electrochemical performances, especially in the first tens of charge/discharge cycles [14,22,23,24,25,26,27,28,29,30,31]. However, the improved electrochemical performances of these composites fade away during cycling because the coatings of carbon or polymer fall off quickly under the erosion of the electrolyte due to the weak coupling between the metal sulfides and carbonaceous matrix.
Nanomaterials possessing unique advantages have been applied widely in sensors, capacitors, solar cells, and especially LIBs. MnS can theoretically deliver a high lithium storage capacity of 616 mAh g−1, as an anode material, while exhibiting a poor cycling performance due to its low electrical conductivity [32]. Some MnS-based composites have been synthesized, aiming to improve conductivity, which exhibited better electrochemical performance than pristine MnS, while their long-term cycling performances were still poor [33,34,35,36]. Moreover, the multistep and complicated procedures of synthesizing these composites limit their large-scale production. Developing a simple, scalable method to prepare transition metal sulfide based nanocomposites is still beneficial.
In this study, the hierarchical MoS2/α-MnS nanocomposites have been synthesized through a one-pot hydrothermal method. These hierarchical MoS2/α-MnS nanocomposites, when being tested as the anode material in LIBs, exhibit a reversible capacity of 1498 mAh g−1 at 50 mA g−1 and deliver a rate capability of 907 mAh g−1 at 1200 mA g−1. The high capacity, 800 mAh g−1, is maintained for over 80 cycles at 300 mA g−1. Moreover, the capacity increase occurs during discharging/charging cycles, which is owed to the metallic nanoparticles that can catalyze the developing polymeric film and enhance conductivity. The excellent electrochemical performances have been discussed in detail and are attributed to the synergistic effect.

2. Materials and Methods

2.1. Materials

All chemicals were used as received without any purification and were analytical grade. Na2MoO4∙H2O, ethanol, CH3CSNH2, KMnO4, H2C2O4, and NaBH4 were supplied by China National Pharmaceutical Group Corporation (SINOPHARM).

2.2. Preparation of MoS2/α-MnS

Hierarchical MoS2/α-MnS nanocomposites were synthesized via a modified hydrothermal route [16]. In a typical synthesis, first 4.7 mmol Na2MoO4∙H2O, 16.7 mmol CH3CSNH2, and 4.6 mmol H2C2O4 were added to 100 mL of DI water; then, a homogeneous, purple solution formed after 2.97 mmol KMnO4 were added under magnetic stirring. The purple solution was transferred to two 50 mL Teflon-lined stainless-steel autoclaves, which were heated at 200 °C for 24 h. Finally, the resultants were naturally cooled with oven, centrifuged, and washed with DI water/ethanol several times. The products were obtained after the washed resultants dried at 60 °C in vacuum for 6 h.

2.3. Electrochemical Measurements

For preparing the working electrodes, active material, conductive graphite, and carboxymethyl celluloses (CMC) were mixed in a weight ratio of 70:20:10; after being ground for 30 min, the mixture was added to deionized water to make homogeneous slurries under stirring; the slurries were cast on copper foils. The coated copper foils were dried under vacuum at 105 °C for 12 h and then transferred to an argon-filled glove box. In this box, CR2016-type coin cells were fabricated, with polyethylene film as separator, pure lithium metal foils as counter electrode, and 1 M LiPF6 dissolved in a mixed solvent containing ethylene carbonate and dimethyl carbonate (1:1 vol%) as electrolyte. Galvanostatic discharge/charge measurements were performed on a Neware battery tester within a voltage window of 3–0.01 V versus Li+/Li. Cyclic voltammetry (CV) test was investigated on an electrochemical workstation (CHI600E, China) in the voltage window of 0.01–2.8 V at a scan rate of 0.1 mV S−1.

2.4. Materials Characterization

The microstructure and morphology of hierarchical MoS2/α-MnS nanocomposites were characterized by scanning electron microscopy (SEM; Hitachi S4800) equipped with an energy dispersive X-ray spectrometer (EDX) and transmission electron microscopy (TEM; JEOL JEM-2010). The crystalline structure was obtained by X-ray powder diffraction (XRD; D/max 2550V, Cu Kα radiation).

3. Results and Discussion

Figure 1 shows the XRD patterns of the hierarchical MoS2/α-MnS nanocomposites. The diffraction peaks located at 14.5°, 32.6°, 39.5°, and 58.2° correspond to the (002), (100), (103), and (110) crystal planes of 2H-MoS2 (JCPDS No. 37-1492) [14,15,16]; The peak at 14.5° corresponds to the (002) plane, indicating 2H-MoS2 has a layered structure [18]. The diffraction peaks at 29.6°, 34.6°, 49.5°, 58.8°, and 61.6° can be assigned to the (111), (200), (220), (311), and (222) crystal planes of α-MnS (JCPDS No. 06-0518), respectively [22,33,34,35,36]. The intensive peaks indicate the good crystal structure of α-MnS.
Figure 2 presents the morphology and microstructure of the hierarchical MoS2/α-MnS nanocomposites. The SEM image (Figure 2a) demonstrates that MoS2/α-MnS has a rose-like hierarchical micromorphology. The TEM images in Figure 2b,c, reveal the microstructure of MoS2/α-MnS, showing MoS2 nanosheets wrapped by α-MnS nanoparticles. The high-resolution TEM image (HRTEM) further provides information about the microstructure (Figure 2d). An interlayer distance of 0.61 nm, corresponding to the (002) crystal plane, indicates the multi-layer nanostructure of MoS2 [37]; a distance of 0.26 nm corresponds to the (200) crystal plane of α-MnS [22,33]. The bright concentric rings and broad diffused halos (shown in the inset of Figure 2d) demonstrate both MoS2 and α-MnS polycrystalline nanostructures. The images of the elemental mapping by energy dispersive spectroscopy (EDS), as shown in Figure 2e, indicate that the elements of Mo, S, and Mn are uniformly distributed. The EDS spectrum of MoS2/α-MnS is depicted in Figure 2f, which confirms that the sample contains elements of Mn, S, and Mo as well as Si from the substrate.
The electrochemical performances of the hierarchical MoS2/α-MnS nanocomposites were utilized as the anode materials (Figure 3). The discharging/charging tests performed at a current density of 200 mA g−1 (Figure 3a). Well-defined plateaus, located at about 2.1, 1.6 and 0.5 V, are observed; the 1st, 2nd, 10th, 50th, 100th, and 150th discharge/charge cycles deliver discharge/charge capacities of 743/580, 587/567, 616/607, 710/700, 889/869, and 959/939 mAh g−1, respectively, and the coulombic efficiencies are 78.1%, 96.6%, 98.5%, 98.6%, 97.8%, and 97.9%, respectively. Note that the capacity of the 150th discharge cycle is nearly 200 mAh g−1 more than that of the 1st discharge cycle, and the capacity keeps rising from the 2nd discharge/charge cycle on until the 150th cycle.
Figure 3b exhibits the cyclic voltammetry (CV) curves of the hierarchical MoS2/α-MnS nanocomposites, further revealing electrochemical reactions during the lithiation/delithiation process. Reduction peaks appear at 2.1, 1.2, 0.6, 0.45, and 0.16 V in the first cycle during the cathodic process. The peak around 2.1 V corresponds to the occurrence of lithium polysulfides [15,16,17,21]; the peak around 1.2 V is assigned to the formation of LixMoS2 through Li+ ions insertion into MoS2; the peak around 0.6 V suggests a part of LixMoS2 further converts into Li2S and metallic Mo nanoparticles [14,15,16,17,18,37]; the peak around 0.45 V can be associated with the further reduction of LixMnS, with Li+ ions resulting in the reduction of Mn2+ to Mn [19,22]; and the peak around 0.16 V corresponds to the developing of a polymeric film [38,39,40]. During the anodic process, three oxidation peaks at 1.3, 1.9, and 2.3 V are observed. The peak at 1.3 V corresponds to the Li+ ions extracted from LixMoS2; the peak at 1.9 V indicates that Li2S reacted with Mn; the peak at 2.3 V denotes that Li2S was oxidized [14,15,16,17,18,37]. Based on the above analysis, the metallic nanoparticles of Mo and Mn that occurred during the charging/discharging process improve the conductivity, which can enhance the high-rate ability of active materials [14,17,19,22]. Moreover, the polymeric film may facilitate hindrance of the dissolution of the polysulfides from the electrode [17,40]. Therefore, for the half cell made of MoS2/α-MnS and Li foil, the possible electrochemical reactions during the discharging/charging process can be described as follows [14,15,16,17,21,33,36,41]:
MnS + xLi + + xe Li x MnS  
Li 2 x MnS + 2 Li + + 2 e Mn + Li 2 S  
MoS 2 + xLi + + xe Li x MoS 2
Li x MoS 2 + ( 4 x ) ( Li + + e ) Mo + 2 Li 2 S
S + 2 ( Li + + e ) Li 2 S  
Figure 3c depicts the cyclic performance of the hierarchical MoS2/α-MnS nanocomposites tested within 0.01–3 V at 300 mA g−1. As shown in Figure 3c, the capacities evidently increase from the 2nd cycle to the 100th cycle, with a capacity increase of 0.32% per cycle, a slight increase from the 101st cycle to the 160th cycle, and a slow decrease from the 170th cycle onward. The capacity increasing with cycles has been observed in other cycling performance tests (Figure S1). This capacity-increase behavior during cycling has been largely attributed to the conductivity being enhanced by the metallic nanoparticles of Mo [15]. The coulombic efficiency in the initial cycle is 79.4%, which then arrives to ~97%, indicating high structural stabilization. Figure 3d shows that the rate capability of the hierarchical MoS2/α-MnS nanocomposites has been evaluated at different current densities within 0.01–3.0 V. The average discharge capacities are around 942, 957, 963, 969, 977, 998, and 907 mAh g−1 at current densities (for 10 cycles each) of 100, 200, 300, 400, 500, 800, and 1200 mA g−1, respectively. When the current density goes back stepwise to 250 and 50 mA g−1, the average discharge capacity arrives to 1136 and 1498 mAh g−1, respectively. The capacity increase also occurs with a stepwise increment of the current density (except for 1200 mA g−1) during rate-capability testing. Some metal oxide based or transition metal sulfide based composites also exhibit the similar behavior of a capacity increase, owing to interfacial Li storage, pseudo-capacitive behavior, or the electrochemical milling effect [19,22,40,42,43,44]. Another view argues that the metallic nanoparticles occurring in the reactions should account for this capacity increase [38,45]. According to the analysis above, we presume that the metallic Mn and Mo nanoparticles account for the capacity increase in the hierarchical MoS2/α-MnS nanocomposites. Figure 3e,f reveal the morphologies of MoS2/α-MnS nanocomposites after being cycled for 1 cycle and 150 cycles, respectively. As shown in these SEM images, the structure of the electrode maintains its integrity on the whole, even after experiencing over 150 cycles, despite cracking, which is in accordance with its good cycling performances.
To verify the presumption above, the electrochemical performances of the hierarchical MoS2/α-MnS nanocomposites have been evaluated by preventing the Mn and Mo nanoparticles from influencing the electrochemical reactions. According to Figure 3b, for the MoS2/α-MnS nanocomposites, the occurrence of the metallic Mn and Mo nanoparticles is inhibited when the testing voltage window is 0.5–2.5 V [40,43]. Figure 4a–c present the galvanostatic electrochemical performances of the hierarchical MoS2/α-MnS nanocomposites tested at 200 mA g−1 within 0.5–2.5 V. Figure 4a displays the specific capacity versus the voltage curves, showing that the specific capacity fades with the number of cycles in the absence of metallic Mn and Mo nanoparticles. Figure 4b presents the differential capacity curves corresponding to the curves of Figure 4a, providing information about the electrochemical reactions within 0.5–2.5 V. Obviously, the curve of Figure 4b resembles that of Figure 3b within 0.5–2.5 V, indicating that they present the same electrochemical behaviors within 0.5–2.5 V. Figure 4c shows the cyclic performance of the hierarchical MoS2/α-MnS nanocomposites tested at 200 mA g−1 within 0.5–2.5 V, indicating that the capacity fades with the number of cycles. Figure 4d shows the discharge capacity versus the voltage curves of Figure 3a within 0.5–2.5 V, which indicates the capacity increases with the cycles in the presence of metallic Mn and Mo nanoparticles (after being discharged to 0.01 V in the first discharging process, Mo nanoparticles occur in the active materials). Therefore, the capacity increase disappears within 0.5–2.5 V, indicating that the metallic Mn and Mo nanoparticles mainly account for the capacity increase during cycling.
The superior electrochemical performances of the hierarchical MoS2/α-MnS nanocomposites is owed to their synergistic effects (Scheme 1). First, the metallic nanoparticles occurring in the electrochemical reactions not only increase the number of paths for electrons and Li+ ions transport but also catalyze the developing polymeric film that lessens the dissolution of the lithium polysulfides, accommodates the massive Li+ ions, and hinders the agglomeration of the metallic nanoparticles [17,19,22,37,40]. Second, the hybrid nanostructure can shorten the diffusion length of the Li+ ions and active materials have been flooded largely by the electrolyte [6,7,9]. Third, the rose-like morphology possesses plenty of void space, which is beneficial in accommodating the Li+ ions [16,46]. Last, the polymeric film and the SEI along with the Mn and Mo metallic nanoparticles can prevent lithium polysulfides from coming out of the electrode, which can enhance the cyclic performance.

4. Conclusions

In summary, a simple, scalable one-pot hydrothermal method has been developed for preparing hierarchical MoS2/α-MnS nanocomposites. As an anode material in LIBs, the hierarchical MoS2/α-MnS nanocomposites deliver a high reversible capacity, a superior cyclic performance, and a high-rate capability. The high reversible capacity is ~1500 mAh g−1 at a current density of 50 mA g−1; the reversible capacity of ~800 mAh g−1 is maintained for over 100 cycles at 300 mA g−1. The superior performances are mainly ascribed to the synergistic effect between MoS2 and α-MnS. Moreover, metallic nanoparticles occur in the electrochemical reactions and lead to the capacity increase during the cycling process. All in all, although the electrochemical mechanism of the hierarchical MoS2/α-MnS nanocomposites needs to be further developed theoretically and experimentally, compared with other widely used strategies, this study proposes a novelty method to improve the electrochemical performances of the anodes in LIBs.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/cryst12121763/s1, Figure S1: Cycling performances of different MoS2/α-MnS nanocomposites tested at different density current: S1 1500, S2 1000, S3 800, S4 600, S5 200 mA g−1.

Author Contributions

Conceptualization, D.-A.Z.; methodology, D.-A.Z.; software, D.-A.Z. and X.Z.; validation, X.Z. and L.Z.; formal analysis, D.-A.Z. and L.Z.; investigation, D.-A.Z.; resources, D.-A.Z. and Q.W.; data curation, D.-A.Z.; writing—original draft preparation, D.-A.Z.; writing—review and editing, D.-A.Z.; visualization, D.-A.Z.; supervision, D.-A.Z.; project administration, D.-A.Z.; funding acquisition, D.-A.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Fundamental Research Funds for the Central Universities (N2005029).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mo, R.W.; Li, F.; Tan, X.Y.; Xu, P.C.; Tao, R.; Shen, G.R.; Lu, X.; Liu, F.; Shen, L.; Xu, B.; et al. High-quality mesoporous graphene particles as high-energy and fast-charging anodes for lithium-ion batteries. Nat. Commun. 2019, 10, 1474. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Bruce, P.G.; Freunberger, S.A.; Hardwick, L.J.; Tarascon, J.M. Li-O2 and Li-S batteries with high energy storage. Nat. Mater. 2012, 11, 19–29. [Google Scholar] [CrossRef] [PubMed]
  3. Manthiram, A.; Chung, S.H.; Zu, C. Lithium-sulfur batteries: Progress and prospects. Adv. Mater. 2015, 27, 1980–2006. [Google Scholar] [CrossRef] [PubMed]
  4. Kim, Y.; Goodenough, J.B. Lithium insertion into transition-metal monosulfides: Tuning the position of the metal 4s band. J. Phys. Chem. C 2008, 112, 15060–15064. [Google Scholar] [CrossRef]
  5. Yan, J.M.; Huang, H.Z.; Zhang, J.; Liu, Z.J.; Yang, Y. A study of novel anode material CoS2 for lithium ion battery. J. Power Sources 2005, 146, 264–269. [Google Scholar] [CrossRef]
  6. Goriparti, S.; Miele, E.; Angelis De, F.; Di Fabrizio, E.; Zaccaria, R.P.; Capiglia, C. Review on recent progress of nanostructured anode materials for li-ion batteries. J. Power Sources 2014, 257, 421–443. [Google Scholar] [CrossRef] [Green Version]
  7. Bruce, P.G.; Scrosati, B.; Tarascon, J.M. Nanomaterials for rechargeable lithium batteries. Angew. Chem. Int. Ed. 2008, 47, 2930–2946. [Google Scholar] [CrossRef]
  8. Yamakawa, N.; Jiang, M.; Grey, C.P. Investigation of the conversion reaction mechanisms for binary copper (II) compounds by solid-state NMR spectroscopy and X-ray diffraction. Chem. Mater. 2009, 21, 3162–3176. [Google Scholar] [CrossRef]
  9. Liao, F.; Światowska, J.; Maurice, V.; Seyeux, A.; Klein, L.H.; Zanna, S.; Marcus, P. Electrochemical lithiation and passivation mechanisms of iron monosulfide thin film as negative electrode material for lithium-ion batteries studied by surface analytical techniques. Appl. Surf. Sci. 2013, 283, 888–899. [Google Scholar] [CrossRef]
  10. Rui, X.H.; Tan, H.T.; Yan, Q.Y. Nanostructured metal sulfides for energy storage. Nanoscale 2014, 6, 9889–9924. [Google Scholar] [CrossRef] [PubMed]
  11. Yu, S.H.; Feng, X.; Zhang, N.; Seok, J.; Abruña, H.D. Understanding conversion-type electrodes for lithium rechargeable batteries. Acc. Chem. Res. 2018, 51, 273–281. [Google Scholar] [CrossRef]
  12. Xu, X.; Liu, W.; Kim, Y.; Cho, J. Nanostructured transition metal sulfides for lithium ion batteries: Progress and challenges. Nano Today 2014, 9, 604–630. [Google Scholar] [CrossRef]
  13. Li, X.F.; Shen, J.F.; Li, N.; Ye, M.N. Fabrication of γ-MnS/rGO composite by facile one-pot solvothermal approach for supercapacitor applications. J. Power Sources 2015, 282, 194–201. [Google Scholar] [CrossRef]
  14. Xiao, J.; Wang, X.J.; Yang, X.Q.; Xun, S.D.; Liu, G.; Koech, P.K.; Liu, J.; Lemmon, J.P. Electrochemically induced high capacity displacement reaction of PEO/MoS2/graphene nanocomposites with lithium. Adv. Funct. Mater. 2011, 21, 2840–2846. [Google Scholar] [CrossRef]
  15. Fang, X.P.; Yu, X.Q.; Liao, S.F.; Shi, Y.F.; Hu, Y.S.; Wang, Z.X.; Stucky, G.D.; Chen, L.Q. Lithium storage performance in ordered mesoporous MoS2 electrode material. Microporous Mesoporous Mater. 2012, 151, 418–423. [Google Scholar] [CrossRef]
  16. Zhang, D.A.; Wang, Q.; Wang, Q.; Sun, J.; Xing, L.L.; Xue, X.Y. High capacity and cyclability of hierarchical MoS2/SnO2 nanocomposites as the cathode of lithium-sulfur battery. Electrochim. Acta 2015, 173, 476–482. [Google Scholar] [CrossRef]
  17. Wang, M.; Li, G.D.; Xu, H.Y.; Qian, Y.T.; Yang, J. Enhanced lithium storage performances of hierarchical hollow MoS2 nanoparticles assembled from nanosheets. ACS Appl. Mater. Interfaces 2013, 5, 1003–1008. [Google Scholar] [CrossRef]
  18. Liu, H.; Su, D.W.; Zhou, R.F.; Sun, B.; Wang, G.X.; Qiao, S.Z. Highly ordered mesoporous MoS2 with expanded spacing of the (002) crystal plane for ultrafast lithium ion storage. Adv. Energy Mater. 2012, 2, 970–975. [Google Scholar] [CrossRef]
  19. Hao, Y.; Chen, C.; Yang, X.Y.; Xiao, G.J.; Zou, B.; Yang, J.W.; Wang, C.L. Studies on intrinsic phase-dependent electrochemical properties of MnS nanocrystals as anodes for lithium-ion batteries. J. Power Sources 2017, 338, 9–16. [Google Scholar] [CrossRef] [Green Version]
  20. Poizot, P.; Laruelle, S.; Grugeon, S.; Tarascon, J.M. Rationalization of the low-potential reactivity of 3d-metal-based inorganic compounds toward Li. J. Electrochem. Soc. 2002, 149, 1212–1217. [Google Scholar] [CrossRef]
  21. Zeng, Z.Y.; Zhang, X.W.; Bustillo, K.; Niu, K.Y.; Gammer, C.; Xu, J.; Zheng, H.M. In situ study of lithiation and delithiation of MoS2 nanosheets using electrochemical liquid cell transmission electron microscopy. Nano Lett. 2015, 15, 5214–5220. [Google Scholar] [CrossRef]
  22. Liu, Y.; Qiao, Y.; Zhang, W.X.; Li, Z.; Hu, X.L.; Yuan, L.X.; Huang, Y.H. Coral-like α-MnS composites with N-doped carbon as anode materials for high-performance lithium-ion batteries. J. Mater. Chem. 2012, 22, 24026–24033. [Google Scholar] [CrossRef]
  23. Liu, B.; Liu, Z.; Li, D.; Guo, P.Q.; Liu, D.Q.; Shang, X.N.; Lv, M.Z.; He, D.Y. Nanoscale alpha-MnS crystallites grown on N-S co-doped rGO as a long-life and high-capacity anode material of Li-ion batteries. Appl. Surf. Sci. 2017, 416, 858–867. [Google Scholar] [CrossRef]
  24. Feng, Y.; Zhang, Y.L.; Wei, Y.Z.; Song, X.Y.; Fub, Y.B.; Battaglia, V.S. A ZnS nanocrystal/reduced graphene oxide composite anode with enhanced electrochemical performances for lithium-ion batteries. Phys. Chem. Chem. Phys. 2016, 18, 30630–30642. [Google Scholar] [CrossRef]
  25. Qu, B.H.; Ji, G.; Ding, B.; Lu, M.H.; Chen, W.X.; Lee, J.Y. Origin of the increased Li+-storage capacity of stacked SnS2/graphene nanocomposite. ChemElectroChem 2015, 2, 1138–1143. [Google Scholar] [CrossRef]
  26. Zhang, J.T.; Yu, L.; Lou, X.W. Embedding CoS2 nanoparticles in N-doped carbon nanotube hollow frameworks for enhanced lithium storage properties. Nano Res. 2017, 10, 4298–4304. [Google Scholar] [CrossRef]
  27. Xia, H.C.; Li, K.X.; Guo, Y.Y.; Guo, J.H.; Xu, Q.; Zhang, J.N. CoS2 nanodots trapped within the graphitic structured N-doped carbon spheres with efficient performances for lithium storage. J. Mater. Chem. A 2018, 6, 7148–7154. [Google Scholar] [CrossRef]
  28. Zhou, Y.L.; Yan, D.; Xu, H.Y.; Yang, J.; Qian, Y.T. Multiwalled carbon nanotube@a-C@Co9S8 nanocomposites: A high-capacity and long-life anode material for advanced lithium ion batteries. Nanoscale 2015, 7, 3520–3525. [Google Scholar] [CrossRef]
  29. Yang, X.; Chan, C.Y.; Xue, H.T.; Xu, J.; Tang, Y.B.; Wang, Q.; Wong, T.L.; Lee, C.S. One-pot synthesis of graphene/In2S3 nanoparticle composites for stable rechargeable lithium ion battery. Crystengcomm 2013, 15, 6578–6584. [Google Scholar] [CrossRef]
  30. Tan, R.; Yang, J.L.; Hu, J.T.; Wang, K.; Zhao, Y.; Pan, F. Core-shell nano-FeS2@N-doped graphene as an advanced cathode material for rechargeable Li-ion batteries. Chem. Commun. 2015, 52, 986–989. [Google Scholar] [CrossRef]
  31. Zhu, J.L.; Li, Y.Y.; Kang, S.; Shen, P.K. One-step synthesis of Ni3S2 nanoparticles wrapped with in situ generated nitrogen-self-doped graphene sheets with highly improved electrochemical properties in Li-ion batteries. J. Mater. Chem. A 2014, 2, 3142–3147. [Google Scholar] [CrossRef]
  32. Cao, X.; Li, H.; He, J.; Kang, L.; Jiang, R.; Shi, F.; Xu, H.; Lei, Z.; Liu, Z. Preparation and formation process of α-MnS@MoS2 microcubes with hierarchical core/shell structure. J. Colloid Interface Sci. 2017, 507, 18–26. [Google Scholar] [CrossRef]
  33. Chen, D.Z.; Quan, H.Y.; Luo, X.B.; Luo, S.L. 3-D graphene cross-linked with mesoporous MnS clusters with high lithium storage capability. Scr. Mater. 2014, 76, 1–4. [Google Scholar] [CrossRef]
  34. Wang, D.D.; Cai, D.P.; Qu, B.H.; Wang, T.H. Rational combination of α-MnS/rGO nanocomposites for high-performance lithium-ion batteries. CrystEngComm 2016, 18, 6200–6204. [Google Scholar] [CrossRef]
  35. Yu, J.; Tang, H. Solvothermal synthesis of novel flower-like manganese sulfide particles. J. Phys. Chem. Solids 2008, 69, 1342–1345. [Google Scholar] [CrossRef]
  36. Biswas, S.; Kar, S.; Chaudhuri, S. Solvothermal synthesis of α-MnS single crystals. J. Cryst. Growth 2005, 284, 129–135. [Google Scholar] [CrossRef]
  37. Chang, K.; Chen, W. In situ synthesis of MoS2/graphene nanosheet composites with extraordinarily high electrochemical performance for lithium ion batteries. Chem. Commun. 2011, 47, 4252–4254. [Google Scholar] [CrossRef]
  38. Laruelle, S.; Grugeon, S.; Poizot, P.; Dollé, M.; Tarascon, J.M. On the origin of the extra electrochemical capacity displayed by MO/Li cells at low potential. J. Electrochem. Soc. 2002, 149, A627–A634. [Google Scholar] [CrossRef]
  39. Kilibarda, G.; Szabó, D.V.; Schlabach, S.; Winkler, V.; Bruns, M.; Hanemann, T. Investigation of the degradation of SnO2 electrodes for use in Li-ion cells. J. Power Sources 2013, 233, 139–147. [Google Scholar] [CrossRef]
  40. Yang, L.C.; Wang, S.N.; Mao, J.J.; Deng, J.W.; Gao, Q.S.; Tang, Y.; Schmidt, O.G. Hierarchical MoS2/polyaniline nanowires with excellent electrochemical performance for lithium-ion batteries. Adv. Mater. 2013, 25, 1180–1184. [Google Scholar] [CrossRef]
  41. Sen, U.K.; Mitra, S. High-rate and high-energy-density lithium-ion battery anode containing 2D MoS2 nanowall and cellulose binder. ACS Appl. Mater. Interfaces 2013, 5, 1240–1247. [Google Scholar] [CrossRef]
  42. Xu, D.; Jiao, R.R.; Sun, Y.W.; Sun, D.Z.; Zhang, X.X.; Zeng, S.Y.; Di, Y.Y. L-cysteine-assisted synthesis of urchin-Like γ-MnS and its lithium storage properties. Nanoscale Res. Lett. 2016, 11, 444–453. [Google Scholar] [CrossRef] [Green Version]
  43. Du, Y.P.; Yin, Z.Y.; Zhu, J.X.; Huang, X.; Wu, X.J.; Zeng, Z.Y.; Yan, Q.Y.; Zhang, H. A general method for the large-scale synthesis of uniform ultrathin metal sulphide nanocrystals. Nat. Commun. 2012, 3, 1177–1183. [Google Scholar] [CrossRef] [Green Version]
  44. Zhang, D.W.; Chen, C.H.; Zhang, J.; Ren, F. Novel electrochemical milling method to fabricate copper nanoparticles and nanofibers. Chem. Mater. 2005, 17, 5242–5245. [Google Scholar] [CrossRef]
  45. Hassan, M.F.; Guo, Z.P.; Chen, Z.X.; Liu, H.K. α-Fe2O3 as an anode material with capacity rise and high rate capability for lithium-ion batteries. Mater. Res. Bull. 2011, 46, 858–864. [Google Scholar] [CrossRef]
  46. Xing, L.L.; He, B.; Nie, Y.X.; Deng, P.; Cui, C.X.; Xue, X.Y. SnO2–MnO2–SnO2 sandwich-structured nanotubes as high-performance anodes of lithium ion battery. Mater. Lett. 2013, 105, 169–172. [Google Scholar] [CrossRef]
Figure 1. XRD of hierarchical MoS2/α-MnS nanocomposites.
Figure 1. XRD of hierarchical MoS2/α-MnS nanocomposites.
Crystals 12 01763 g001
Figure 2. (a) SEM; (b) and (c) TEM at different magnifications; (d) HRTEM; (e) EDS mapping for elements of Mn, S, and Mo; (f) EDS spectrum of hierarchical MoS2/α-MnS nanocomposites.
Figure 2. (a) SEM; (b) and (c) TEM at different magnifications; (d) HRTEM; (e) EDS mapping for elements of Mn, S, and Mo; (f) EDS spectrum of hierarchical MoS2/α-MnS nanocomposites.
Crystals 12 01763 g002
Figure 3. The galvanostatic measurements tested in the voltage range of 0.01–3.0 V for hierarchical MoS2/α−MnS nanocomposites. (a) Discharge–charge profiles at 200 mA g−1. (b) CV curves tested in the voltage window of 0.01−2.8 V (versus Li+/Li) at a scan rate of 0.1 mV s−1. (c) Cycling−performance profiles at 300 mA g−1. (d) Rate−capability profiles at different rates. (e,f) SEM image electrode after 1 cycle and 150 cycles, respectively.
Figure 3. The galvanostatic measurements tested in the voltage range of 0.01–3.0 V for hierarchical MoS2/α−MnS nanocomposites. (a) Discharge–charge profiles at 200 mA g−1. (b) CV curves tested in the voltage window of 0.01−2.8 V (versus Li+/Li) at a scan rate of 0.1 mV s−1. (c) Cycling−performance profiles at 300 mA g−1. (d) Rate−capability profiles at different rates. (e,f) SEM image electrode after 1 cycle and 150 cycles, respectively.
Crystals 12 01763 g003
Figure 4. The galvanostatic measurements tested in the voltage range of 0.5−2.5 V for hierarchical MoS2/α−MnS nanocomposites. (a) Discharge−charge profiles at 200 mA g−1. (b) The corresponding derivative relation of the galvanostatic voltage versus the capacity. (c) Cycling performance profiles at 200 mA g−1. (d) Discharge capacity versus voltage curves of Figure 3a within 0.5−2.5 V.
Figure 4. The galvanostatic measurements tested in the voltage range of 0.5−2.5 V for hierarchical MoS2/α−MnS nanocomposites. (a) Discharge−charge profiles at 200 mA g−1. (b) The corresponding derivative relation of the galvanostatic voltage versus the capacity. (c) Cycling performance profiles at 200 mA g−1. (d) Discharge capacity versus voltage curves of Figure 3a within 0.5−2.5 V.
Crystals 12 01763 g004
Scheme 1. The lithiation/delithiation of MoS2/α-MnS nanocomposites.
Scheme 1. The lithiation/delithiation of MoS2/α-MnS nanocomposites.
Crystals 12 01763 sch001
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, D.-A.; Zhao, X.; Zhu, L.; Wang, Q. Facile One-Pot Hydrothermal Synthesis of Hierarchical MoS2/α-MnS Nanocomposites with Good Cycling Performance as Anode Materials for Lithium-Ion Batteries. Crystals 2022, 12, 1763. https://doi.org/10.3390/cryst12121763

AMA Style

Zhang D-A, Zhao X, Zhu L, Wang Q. Facile One-Pot Hydrothermal Synthesis of Hierarchical MoS2/α-MnS Nanocomposites with Good Cycling Performance as Anode Materials for Lithium-Ion Batteries. Crystals. 2022; 12(12):1763. https://doi.org/10.3390/cryst12121763

Chicago/Turabian Style

Zhang, De-An, Xishan Zhao, Lin Zhu, and Qi Wang. 2022. "Facile One-Pot Hydrothermal Synthesis of Hierarchical MoS2/α-MnS Nanocomposites with Good Cycling Performance as Anode Materials for Lithium-Ion Batteries" Crystals 12, no. 12: 1763. https://doi.org/10.3390/cryst12121763

APA Style

Zhang, D.-A., Zhao, X., Zhu, L., & Wang, Q. (2022). Facile One-Pot Hydrothermal Synthesis of Hierarchical MoS2/α-MnS Nanocomposites with Good Cycling Performance as Anode Materials for Lithium-Ion Batteries. Crystals, 12(12), 1763. https://doi.org/10.3390/cryst12121763

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