Next Article in Journal
Sustainable Myrcene-Based Elastomers via a Convenient Anionic Polymerization
Next Article in Special Issue
Facile Synthesis of Carboxymethyl Cellulose Coated Core/Shell SiO2@Cu Nanoparticles and Their Antifungal Activity against Phytophthora capsici
Previous Article in Journal
Hybrid Filaments from Saccaharina lattisima Biomass: Engineering of Alginate Properties with Maleic Anhydride Grafted Linseed Oil
Previous Article in Special Issue
Gemcitabine-Loaded Magnetically Responsive Poly(ε-caprolactone) Nanoparticles against Breast Cancer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Glucose-Assisted One-Pot Hydrothermal Synthesis of Hierarchical-Structured MoS2/C Quasi-Hollow Microspheres for High-Performance Lithium Ion Battery

State Key Laboratory of Polymer Materials Engineering, Polymer Research Institute, Sichuan University, Chengdu 610065, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Polymers 2021, 13(5), 837; https://doi.org/10.3390/polym13050837
Submission received: 16 January 2021 / Revised: 25 February 2021 / Accepted: 25 February 2021 / Published: 9 March 2021
(This article belongs to the Special Issue Nanoparticle Functionalization by Polymers: Methods and Applications)

Abstract

:
In this work, hierarchical MoS2/C quasi-hollow microspheres are prepared by a one-pot hydrothermal process with the addition of glucose. The glucose is not only inclined to form the roundish sphere in the completion of the synthesis of MoS2, but at the same time the microspheres formed by the glucose can act as the nuclei on which the MoS2 grows. Glucose, acting as a nucleating agent, has the advantages of being low-cost and environmentally friendly, which can simplify the fabrication process. The interiors of the MoS2/C samples are multi-hole and quasi-hollow, which is beneficial for the insertion and extraction of lithium ions. For the first time, we demonstrate that hierarchical-structured MoS2/C quasi-hollow microspheres exhibit an excellent cycling stability and rate capability in lithium ion batteries (LIBs) and are significantly superior to the bulk MoS2. The method presented in this article may provide a simple, clean. and economical strategy for the preparation of MoS2/C microspheres as a feasible and promising anode material for LIBs.

1. Introduction

New energy technology is considered to be one of the five major technical areas in the development of the 21st century world economy, which receives great attention around the world. Rechargeable lithium ion batteries (LIBs) are currently key components of portable, entertainment, computing, and telecommunication equipment due to their light weight, high specific capacity, low self-discharge rate, good cycling performance, having no memory effect, and being eco-friendly [1,2]. Due to the limited lithium storage capacity of the commercial graphite anode (372 mAh g−1) in LIBs, they can no longer suit the fast growth of the modern society. Therefore, the need to explore alternative anode materials with a smaller size, lighter weight, and higher reversible capacity is more demanding than ever [3,4].
As a typical type of layered transition metal dichalcogenide (TMD), MoS2 has a sandwich structure, stacked by S–Mo–S tri-layers through Van Der Waals interaction [5,6]. This unique structure endows MoS2 with numerous applications in the fields of lubricants [7], catalytic hydrodesulfuration [8], hydrogen storage [9], evolution reactions [10,11], and anode materials for Li+/Na+/Mg2+ batteries [12,13,14,15,16,17,18]. Following the first patent of MoS2 as the anode material in LIBs, a large number of MoS2 with different morphologies (such as nanoflake, nanoflower, hierarchical microsphere) were fabricated for LIBs [5,19,20,21]. As an anode material, MoS2 has a specific capacity of up to 1000 mAh g-1 [22], but its cycling stability is unsatisfactory due to the great volume expansion-induced mechanical strain upon the cycling, which subsequently causes electrode pulverization and the loss of active material/current collector integrity [6,23,24]. On the other hand, the relatively low electric conductivity of bulk MoS2 also poses a negative effect on its rate capability in LIBs.
To address the above issues, two approaches have been proposed [25]. One is to devise special nanostructures of MoS2 (such as nanotubes [26], nanoflakes [27], and mesoporous [28] and hierarchical micro-nanostructures [29,30]) so that they can release the mechanical strain of the volume change and lessen the barrier for the Li+ transport. However, the aggregation of nano-structured MoS2 may still happen during the charge/discharge process and it will still suffer from a low electric conductivity. The other approach is to fabricate the MoS2 composites with some electrically conductive carbons or polymers, which obviously is beneficial in terms of improving the electrical conductivity of the electrode material [31,32,33,34,35,36,37,38]. Therefore, the combination of these two strategies—i.e., the preparation of a MoS2/electrically conductive carbon composite, in which MoS2 presents a special morphology—may offer a perfect solution to the problems of bulk MoS2, as mentioned above. Recently, hierarchical hollow nanoparticles have aroused great interest because they not only effectively increase the electrolyte/electrode contact area and shorten the pathway of ion diffusion, but also buffer the volume expansion of electrodes, thus stabilizing the electrode structure under the charge–discharge process. Although MoS2/C hollow microspheres have been reported once [25,29], the involvement of mono dispersed polystyrene microspheres as templates makes the synthetic process expensive and sophisticated.
Glucose, a widely accessible biochemical, is a source of carbon, which is also considered a method that adds value to biorefinery. In this paper, we present a fast, simple, and low-cost method to prepare MoS2/C quasi-hollow microspheres using a one-pot hydrothermal method with the assistance of glucose. Compared with other hydrothermal methods to prepare molybdenum disulfide/carbon composites [34,35], we selected ammonium molybdate and thiourea as the raw materials to synthesize MoS2 at the molecular level, which can make the glucose molecules first aggregate into nuclei and then cause ammonium molybdate and thiourea to grow in situ on the surface. First, the MoS2 layer grows through heterogeneous nucleation on the surface of the spherical core formed by glucose aggregation. Second, the glucose is pyrolyzed into porous carbon in the interior to form the MoS2/C microspheres with a quasi-hollow structure after heat treatment. As expected, the MoS2/C quasi-hollow microspheres can not only improve the cycling stability of MoS2, but also enhance the electric conductivity of the electrode material, resulting in an excellent rate performance in LIBs.

2. Experimental Section

2.1. Synthesis of Hierarchical MoS2/C Quasi-Hollow Microspheres

Ammonium molybdate tetrahydrate ((NH4)6MoVI7O24·4H2O, 99%) and thiourea (C(NH2)2S, 99%) were purchased from Shanghai Aladdin Co., Ltd. Glucose was purchased from Kelong Chemical Co., Ltd. (Chengdu, China). All the reagents were analytical grade and directly used as received without further purification.
In a typical procedure, 8.6 mmol/L of ammonium molybdate, 140 mmol/L of thiourea, and 148 mmol/l of glucose were dissolved in 30 mL of distilled water. After being stirred for 30 min, the solution was then transferred into a Teflon-lined stainless steel autoclave. The autoclave was sealed and maintained at 240 °C for 48 h. The black precipitates were collected by filtering, washing with distilled water, and freeze-drying for 24 h. Subsequently, the as-prepared black powder was fabricated into MoS2/C composites by annealing in an Ar atmosphere under 800 °C for 2 h. The content of glucose varied between 148 and 296 mmol/L (MoS2/C composites obtained with the addition of 148, 222, and 296 mmol/L of glucose) and such obtained composites are denoted as MoS2/C-1, MoS2/C-2, and MoS2/C-3, respectively, according to the approximate mass ratio of glucose and the total weight of the ammonium molybdate and thiourea. The bulk MoS2 was prepared through the same procedure without the addition of glucose for comparison.

2.2. Characterizations

Powder X-ray diffraction (XRD) patterns were recorded on a Rigaku (Smart lab III) of Hitachi Company (Tokyo, Japan) using Cu Kα radiation within 2θ = 10°–80° with a scanning rate of 10°/min. Product morphologies were observed by field emission scanning electron microscopy (FESEM) and high-resolution transmission electron microscopy (TEM) (Philips-FEI company Hillsboro, OR, USA). X-ray photoelectron spectra (XPS) were collected on an Escalab 250Xi X-ray photoelectron spectrometer (Thermo Fisher Scientific Co., Ltd., Waltham, Ma, USA).

2.3. Electrochemical Measurements

The anodes were fabricated by mixing active materials (MoS2, MoS2/C composites) with acetylene black carbon and polyvinylidene fluoride (PVDF), at a weight ratio of 8:1:1 in the solvent of N-methyl-2-pyrrolidone (NMP). The obtained slurry was cast onto a Cu foil with a wet film thickness of about 150 μm and dried at 120 °C in a vacuum for 12 h. The electrodes were punched to φ=13 mm discs before use. Subsequently, CR 2032 coin-type cells were assembled in an argon-filled glove box using MoS2, or MoS2/C composites as the working electrode, Li foil as the counter electrode, a glass fiber membrane (Whatman 1825-025) as a separator, and 1M of LiPF6 dissolved in a 1:1 mixture of ethylene carbonate and diethyl carbonate as the electrolyte. The cells were measured using an automatic battery tester system of Wuhan Landian Electronics Co., Ltd. (Land CT2001A, Wuhan, China) and galvanostatic charged and discharged at various current densities in the voltage range of 0.01–3.0 V. The cyclic voltammetry (CV) was carried out in the potential range of 0.01–3.0 V (vs. Li/Li+) with a scan rate of 0.2 mV/s, and the electrochemical impedance spectroscopy (EIS) was measured by applying an AC voltage of 10 mV in the frequency range of 100 kHz to 0.01 Hz using the Biologic VMP3 electrochemical workstation.

3. Results and Discussion

3.1. Characterization of Structure and Morphology

Figure 1 is the schematic illustration of the formation of hierarchical MoS2/C quasi-hollow microspheres. First, glucose spontaneously aggregates, acting as a spherical nuclei onto which MoS2, synthesized by the reaction of (NH4)6MoVI7O24·4H2O (Mo source) and C(NH2)2S (S source), grows through heterogeneous nucleation. In the process of hydrothermal reaction, glucose molecules undergo dehydration and cross-linking reaction, carbonize to form carbon-carbon bonds, and gradually form carbon microspheres. Subsequently, after the high-temperature annealing treatment, its aromaticity is further increased, which is beneficial to the improvement of conductivity [38,39]. Then, MoS2/C quasi-hollow microspheres with hierarchical structures are obtained.
Figure 2 shows the XRD pattern of pristine MoS2 and MoS2/C composites prepared with different amounts of glucose. The XRD pattern of the annealed bulk MoS2 can be indexed as the hexagonal crystalized structure of MoS2 (JCPDS no. 37-1492), and the strong (002) peak centered at 2θ = 14.2° indicates a well-stacked layered structure. However, in the MoS2/C composites with a different carbon content, the (002) plane peak disappears, indicating that the stacking layered structure of MoS2 is inhibited in the MoS2/C composites 38. The other two peaks at 2θ = 33.0° and 58.9° are in good agreement with those of pristine MoS2. The XRD results clearly reveal that the structure of MoS2 is mono or has few layers, which is further confirmed by the results of the high-resolution TEM (Figure 3). Besides these, another three weak peaks which do not belong to the MoS2 can also be found in the MoS2/C composites. The weak peak at 2θ = 25° should be attributed to the (002) plane of amorphous carbon, as the annealing temperature of 800 °C is much lower than the graphitization temperature of 3000 °C. In addition, another two weak peaks, marked with * and #, do not belong to the MoS2 or the carbon. According to the Bragg equation, the d-spacing of peak * is 0.97∼1.18 nm and that of peak # is 0.49∼0.58 nm, which do not agree with the d-spacing of MoS2 (0.62 nm) or carbon (0.34 nm). Referring to the MoS2/amorphous carbon composites synthesized by a hydrothermal route [17,38,40,41], it is known that the d-spacing of peak * is very close to the distance of adjacent MoS2 nanosheets in amorphous carbon, suggesting that some amorphous carbon is inserted into the layer of MoS2. The d-spacing of # (0.49∼0.58 nm) can be indexed to the distance between the MoS2 layer and the carbon layer. Besides these peaks, no other peaks are formed in the MoS2/C composites, indicating that impurity does not exist in the MoS2/C composites.
The morphology of the as-prepared pristine MoS2 and MoS2/C composites were first observed by Scanning electron microscopy (SEM). As shown in Figure 3a, MoS2 tend to form a flower-like agglomeration assembled by the MoS2 nanosheets. With the introduction of carbon, the MoS2/C composite shows a hierarchical structure, as shown in Figure 3c–h, which is composed of MoS2 growing on the surface of the C microspheres. This hierarchical structure enlarges the contact area of the anode of the MoS2/C and the electrolyte. Additionally, it can shorten the Li+ diffusion channel in the charge–discharge process. Compared with the pristine MoS2, there was no agglomeration of the MoS2 nanosheets in the MoS2/C hierarchical structure, suggesting that the amorphous carbon plays a vital role in stabilizing the MoS2 nanosheets. At the same time, it can be observed that the MoS2/C quasi-hollow microspheres become bigger and more roundish rather than flocking together by increasing the content of amorphous carbon by varying the amount of glucose in the hydrothermal process. By increasing the amount of glucose, it is easier for glucose to aggregate to individually bigger spherical nuclei through the interaction of oxygen-containing functional groups, as do the resulting MoS2/C microspheres formed by means of heterogeneous nucleation on glucose.
To further investigate the hierarchical structures of the MoS2/C microspheres, TEM characterization was performed on one of the composites of MoS2/C-2. As shown in Figure 4a–d, it is obvious that MoS2/C microspheres exhibit a quasi-hollow structure, the center of which is different from its margin: the interior is mainly composed of porous residual amorphous carbon due to the incomplete decomposition of glucose under Ar, while the bulk shell indicates the existence of the MoS2. The high-resolution TEM images of the MoS2/C microspheres (Figure 4c) reveals that the shells are indeed composed of 1–3 layers of MoS2 and the interlayered spacing is about 0.65 nm which is very close to the value calculated from the XRD patterns by using the Bragg equation. A selected area electron diffraction (SAED) pattern of the quasi-hollow MoS2/C composite displays two bright diffraction rings belonging to the (100) and (110) reflections of the hexagonal-phase MoS2, which is consistent with the XRD results.
X-ray photoelectron spectroscopy (XPS) was conducted to determine the chemical composition of the MoS2/C composites. The atomic ratio of the Mo and S elements calculated from the XPS spectra is 2.02, approaching the stoichiometric value of MoS2 [42]. The Mo 3d XPS spectrum of the MoS2/C-2 sample shows two broad peaks at 229.3 eV and 232.4 eV, which can be assigned to the doublet Mo 3d5/2 and Mo 3d3/2, respectively (Figure 5a) [42,43,44]. The S 2p spectrum can be deconvoluted into two peaks centered at 162.3 eV and 163.5 eV, which correspond to S 2p3/2 and S 2p1/2, respectively (Figure 5b) [42,43,44]. Furthermore, the Mo 3d XPS spectrum and S 2p XPS spectrum suggest that Mo4+ and S2- are the dominant states in the MoS2/C samples [34].

3.2. Electrochemical Performance Characterization

Pristine MoS2 and MoS2/C composites were assembled as a half battery in order to compare their electrochemical properties. Figure 6a–d show the cyclic voltammograms (CV) of pristine MoS2 and MoS2/C composites at a scan rate of 0.2 mV/s. The CV curves of pristine MoS2 show two reductive peaks at 0.8 and 0.3 V and two corresponding oxidation peaks at 1.7 V and 2.2 V in the first cycle. The two reductive peaks at 0.8 V and 0.3 V are due to the reactions (1) and (2) as follows, respectively:
MoS 2 + xLi + + xe Li x MoS 2
Li x MoS 2 + 4 x Li + + 4 x e Mo + 2 Li 2 S
The pronounced oxidation peaks at 1.7 V and 2.2 V could be attributed to the delithiation of Li2S [43,44]. In the second and third cathodic sweeps, a new reductive peak at ∼1.8 V appears, which could be attributed to the following reaction (3):
2 Li + + S + 2 e Li 2 S
In contrast, MoS2/C composites present a similar but slightly different CV behavior compared to the pristine MoS2. During the first anodic scan, there are two peaks located at 0.8 V and 0.6 V in the MoS2/C-1 composite, which could be attributed to the reactions (1) and (2) [41,45]. Compared with MoS2/C-1, the reductive peak of the MoS2/C-2 and MoS2/C-3 at 0.8 V becomes inconspicuous for the existence of carbon interlayers between the neighboring MoS2 layers preventing the phase transformation of MoS2 from trigonal prismatic to octahedral [41]. During the subsequent cathodic scan of the MoS2/C-1 electrode, two oxidative peaks at 1.2 V and 2.2 V are observed, which can be attributed to the oxidation of Li2S into S [6,25,46]. After the first cycle, the electrode is mainly composed of Mo and S instead of the initial MoS2 [6]. In the subsequent scans, the two peaks observed in the first anodic scan disappears. Instead, a new prominent reductive peak at 2.0 V presents, which could be also attributed to the reaction (3) [6,25]. In contrast, in the second scan the anodic peaks at 1.4 V and 2.2 V overlap with those in the first scan, suggesting excellent electrochemical reversibility and stability. The 2.0/2.2 V redox pair constitutes a reversible redox couple, resembling those of sulfur electrodes closely [25].
The galvanostatic discharge/charge profiles of pristine MoS2 and MoS2/C composites are presented in Figure 6e–h, respectively. As shown in Figure 6e, pristine MoS2 presents two potential plateaus at ∼0.3 V and 0.8 V in the first discharge process due to the Li+ intercalation reaction followed by a series of conversion reactions. In contrast, there are two inconspicuous potential plateaus at 0.8 V and 0.6 V on the discharge curves of the MoS2/C-1 composites, which is caused by the conversion reaction of the Li+ intercalation of the MoS2. Nevertheless, only one potential plateau at ∼0.6 V presents at the first discharge curve and the plateau corresponding to the Li+ intercalation disappears. During the subsequent charge process of the pristine MoS2 and MoS2/C composites, a series of potential plateaus are all attributed to the oxidation of Li2S into S. The information reflected by these charge/discharge curves are in agreement with the results of the CV curves. The initial discharge capacities of pristine MoS2, MoS2/C-1, MoS2/C-2, and MoS2/C-3 are 983.6, 867, 924.6, and 771.3 mAh/g, respectively. After 150 continuous charge–discharge cycles, the reversible discharge capacities of pristine MoS2, MoS2/C-1, MoS2/C-2, and MoS2/C-3 retain about 134.8, 467.8, 549.7, and 387.9 mAh/g, respectively, suggesting that the electrochemistry performance of the MoS2/C composites is superior to that of pristine MoS2 and the structural stability of MoS2/C-2 is superior among the MoS2/C composites.
The cycling performances of MoS2/C quasi-hollow microspheres are compared in Figure 7. As shown in Figure 7a, the pristine MoS2 exhibits an initial discharge capacity of 979.2 mAh g-1, which decays quickly within five cycles to 792.2 mAh g−1, then after a short plateau of up to 25 cycles continuously drops, and it only can maintain about 100 mAh g-1 beyond 75 cycles. However, in contrast, the MoS2/C composite electrodes exhibit a significantly improved cycling stability. The initial discharge capacity of the MoS2/C electrodes are 867-, 924.6- and 771.3 mAh g-1 respectively and after a slight cycle loss in the first five cycles, these electrodes remain with reversible capacities of 456.9, 560.5, and 338.8 mAh g-1, respectively, even after 200 cycles. One has to note that the calculated capacity of the composites in which about 62.55%~72.7% of the carbon has no capacity contribution. To make a further comparison, we summarized the electrochemical properties of some of the MoS2/C electrodes as shown in Table 1. It can be seen that the reversible specific capacity obtained from the MoS2/C-2 is comparable to those of the other kinds of MoS2/C electrodes.
Moreover, the MoS2/C-2 electrode demonstrates a better rate performance in all the electrodes as shown in Figure 7b. The capacity of MoS2/C-2 delivered 586, 506, 410, 350, and 300 mAh g-1 at 0.1, 0.2, 0.5, 1, and 2 A g-1, respectively. When the current density recovers to 100 mA g-1 after cycling under the high current densities, MoS2/C-2 sample can still regain a reversible capacity of 550 mA g-1. However, the capacity of the bulk MoS2 fades continually under the current density of 1 A g-1 and drops to ~ 100 mAh g-1 at 2 A g−1. When the current density recovers to 100 mA g-1, the capacity of the bulk MoS2 still drops continually, to far below its initial capacity. The superior electrochemical performance of the MoS2/C electrode can be attributed to the following reasons. Firstly, the MoS2/C quasi-hollow microspheres with a larger specific area can provide a contact area between the electrode material and the electrolyte, which may facilitate the Li+ insertion/distraction in the charge/discharge process. Secondly, the interstices between the ultrathin MoS2 nanosheets and quasi-hollow interior of the MoS2 microspheres play the role of buffering the mechanical stresses induced by the volumetric expansion/shrinkage of the MoS2 [25]. Thirdly, the porous carbon cores may serve as adsorbents to restrain the discharge products of the sulfur (lithium polysulfide) from dissolving into electrolytes [25,47]. Lastly, the excessive carbon is not beneficial for the capacity of the MoS2/C microspheres but favorable to the cycling stability because the carbon in the MoS2/C microspheres is amorphous and it will lower the capacity of the MoS2/C microspheres.
To gain an in-depth understanding of the good electrochemical performance of the MoS2/C sample, the electrochemical impedance spectra (EIS) of the bulk MoS2 and the MoS2/C electrodes before cycling and after cycling 50 cycles are measured and presented in Figure 8, respectively. Two overlapped depressed semi-circles at high frequency along with an inclined spike at low frequency are observed for all the spectra. The two semi-circles could be attributed to the charge–transfer resistance (Rct) between the active material and the electrolyte, and correspond to the lithium diffusion process within the electrodes [38,48,49]. The charge–transfer resistance of the bulk MoS2, MoS2/C-1, MoS2/C-2, and MoS2/C-3 is approximately 110, 98, 95, and 60 Ω before cycling, and the corresponding values are about 120, 80, 37, and 50 Ω after the 50th cycle of the charge–discharge process, respectively. It can be seen that the MoS2/C-2 composite has a much lower charge–transfer resistance among all the electrodes, which is beneficial for the transmission of Li+ and electrons during the charge–discharge process. Furthermore, the much lower impedance of the MoS2/C composites demonstrates that the existence of porous carbon can greatly enhance the conductivity of the MoS2/C electrode.

4. Conclusions

In summary, hierarchical quasi-hollow MoS2/C microspheres were successfully prepared by a facile one-pot approach with the assistance of glucose. The as-prepared MoS2/C microspheres are significant due to their hierarchical structures integrated with the MoS2 nanostructure, in which the aggregation of MoS2 is effectively prevented for the porous carbon microspheres. These unique hierarchical quasi-hollow MoS2/C microspheres exhibited a greatly improved Li-ion storage performance. In particular, the MoS2/C composites exhibited outstanding cycling stability and better rate performance than the bulk MoS2. Among all the electrodes, the MoS2/C-2 electrode exhibited a higher capacity and better rate performance than the other electrodes. The amorphous carbon in the interior of the MoS2/C microspheres can prevent the aggregation of MoS2 in the discharge/charge process and promote the conductivity of the MoS2/C microspheres simultaneously. However, the excessive amorphous carbon will result in a decrease in the capacity. It can be believed that the facile and low-cost synthesis of quasi-hollow MoS2/C microspheres with a better performance has promising applications in the energy storage and conversion system.

Author Contributions

X.L. (Xingang Liu) and J.T. (Jiang Tan) conceived and designed the experiments; J.T. (Jiang Tan) and X.L. (Xi Li) performed the experiments and analyzed the data; X.L. (Xingang Liu) wrote the paper. Supervision, project administration and funding acquisition, C.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Key R&D Program of China (no. 2017YFE0111500) and the National Natural Science Foundation of China (nos. 51933007 and 51673123).

Institutional Review Board Statement

Not applicable.

Conflicts of Interest

The authors declare no competing financial interest.

References

  1. Fan, Y.C.; Chen, X.; Legut, D.; Zhang, Q.F. Modeling and theoretical design of next-generation lithium metal batteries. Energy Storage Mater. 2019, 16, 169–193. [Google Scholar] [CrossRef]
  2. Winter, M.; Barnett, B.; Xu, K. Before Li Ion Batteries. Chem. Rev. 2018, 118, 11433–11456. [Google Scholar] [CrossRef]
  3. Zhang, X.Q.; Zhao, C.Z.; Huang, J.Q.; Zhang, Q. Recent Advances in Energy Chemical Engineering of Next-Generation Lithium Batteries. Engineering 2018, 4, 831–847. [Google Scholar] [CrossRef]
  4. Hewathilake, H.; Karunarathne, N.; Wijayasinghe, A.; Balasooriya, N.W.B.; Arof, A.K. Performance of developed natural vein graphite as the anode material of rechargeable lithium ion batteries. Ionics 2017, 23, 1417–1422. [Google Scholar] [CrossRef]
  5. Liu, Y.; Zhao, Y.; Jiao, L.; Chen, J. A graphene-like MoS2/graphene nanocomposite as a highperformance anode for lithium ion batteries. J. Mater. Chem. A 2014, 2, 13109–13115. [Google Scholar] [CrossRef]
  6. Yang, L.; Wang, S.; Mao, J.; Deng, J.; Gao, Q.; 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]
  7. Li, H.; Wang, J.H.; Gao, S.; Chen, Q.; Peng, L.M.; Liu, K.H.; Wei, X.L. Superlubricity between MoS2 Monolayers. Adv. Mater. 2017, 29, 1701474. [Google Scholar] [CrossRef]
  8. Wang, Y.; Mao, J.; Meng, X.G.; Yu, L.; Deng, D.H.; Bao, X.H. Catalysis with Two-Dimensional Materials Confining Single Atoms: Concept, Design, and Applications. Chem. Rev. 2019, 119, 1806–1854. [Google Scholar] [CrossRef]
  9. Eftekhari, A. Electrocatalysts for hydrogen evolution reaction. Int. J. Hydrogen Energy 2017, 42, 11053–11077. [Google Scholar] [CrossRef]
  10. Shi, Y.; Zhou, Y.; Yang, D.R.; Xu, W.X.; Wang, C.; Wang, F.B.; Xu, J.J.; Xia, X.H.; Chen, H.Y. Energy Level Engineering of MoS2 by Transition-Metal Doping for Accelerating Hydrogen Evolution Reaction. J. Am. Chem. Soc. 2017, 139, 15479–15485. [Google Scholar] [CrossRef]
  11. Merki, D.; Hu, X. Recent developments of molybdenum and tungsten sulfides as hydrogen evolution catalysts. Energy Environ. Sci. 2011, 4, 3878–3888. [Google Scholar] [CrossRef] [Green Version]
  12. Wang, H.Q.; Pan, Q.C.; Wu, Q.; Zhang, X.H.; Huang, Y.G.; Lushington, A.; Li, Q.Y.; Sun, X.L. Ultrasmall MoS2 embedded in carbon nanosheets-coated Sn/SnOx as anode material for high-rate and long life Li-ion batteries. J. Mater. Chem. A 2017, 5, 4576–4582. [Google Scholar] [CrossRef]
  13. Sun, D.; Ye, D.L.; Liu, P.; Tang, Y.G.; Guo, J.; Wang, L.Z.; Wang, H.Y. MoS2/Graphene Nanosheets from Commercial Bulky MoS2 and Graphite as Anode Materials for High Rate Sodium-Ion Batteries. Adv. Energy Mater. 2018, 8, 11. [Google Scholar] [CrossRef]
  14. Liu, Y.C.; Fan, L.Z.; Jiao, L.F. Graphene intercalated in graphene-like MoS2: A promising cathode for rechargeable Mg batteries. J. Power Sources 2017, 340, 104–110. [Google Scholar] [CrossRef]
  15. Ding, S.; Chen, J.S.; Lou, X.W. Glucose-Assisted Growth of MoS2 Nanosheets on CNT Backbone for Improved Lithium Storage Properties. Chem. A Eur. J. 2011, 17, 13142–13145. [Google Scholar] [CrossRef]
  16. Liang, Y.; Feng, R.; Yang, S.; Ma, H.; Liang, J.; Chen, J. Rechargeable Mg Batteries with Graphene-like MoS2 Cathode and Ultrasmall Mg Nanoparticle Anode. Adv. Mater. 2011, 23, 640–643. [Google Scholar] [CrossRef]
  17. Liu, Y.; Jiao, L.; Wu, Q.; Du, J.; Zhao, Y.; Si, Y.; Wang, Y.; Yuan, H. Sandwich-structured graphene-like MoS2/C microspheres for rechargeable Mg batteries. J. Mater. Chem. A 2013, 1, 5822–5826. [Google Scholar] [CrossRef]
  18. Choi, S.H.; Ko, Y.N.; Lee, J.-K.; Kang, Y.C. 3D MoS2-Graphene Microspheres Consisting of Multiple Nanospheres with Superior Sodium Ion Storage Properties. Adv. Funct. Mater. 2015, 25, 1780–1788. [Google Scholar] [CrossRef]
  19. Nguyen, T.M.N.; Vuong, V.D.; Phong, T.M.; Le, T.V. Fabrication of MoS2 Nanoflakes Supported on Carbon Nanotubes for High Performance Anode in Lithium-Ion Batteries (LIBs). J. Nanomater. 2019, 2019, 7. [Google Scholar] [CrossRef]
  20. Ette, P.M.; Chithambararaj, A.; Prakash, A.S.; Ramesha, K. MoS2 Nanoflower-Derived Interconnected CoMoO4 Nanoarchitectures as a Stable and High Rate Performing Anode for Lithium-Ion Battery Applications. ACS Appl. Mater. Interfaces 2020, 12, 11511–11521. [Google Scholar] [CrossRef]
  21. Tang, W.J.; Wang, X.L.; Zhong, Y.; Xie, D.; Zhang, X.Q.; Xia, X.H.; Wu, J.B.; Gu, C.D.; Tu, J.P. Hierarchical MoS2/Carbon Composite Microspheres as Advanced Anodes for Lithium/Sodium-Ion Batteries. Chem. A Eur. J. 2018, 24, 11220–11226. [Google Scholar] [CrossRef]
  22. Feng, C.; Ma, J.; Li, H.; Zeng, R.; Guo, Z.; Liu, H. Synthesis of molybdenum disulfide (MoS2) for lithium ion battery applications. Mater. Res. Bull. 2009, 44, 1811–1815. [Google Scholar] [CrossRef]
  23. Hu, Y.Y.; Bai, Y.L.; Wu, X.Y.; Wei, X.; Wang, K.X.; Chen, J.S. MoS2 nanoflakes integrated in a 3D carbon framework for high-performance sodium-ion batteries. J. Alloy. Compd. 2019, 797, 1126–1132. [Google Scholar] [CrossRef]
  24. Stephenson, T.; Li, Z.; Olsen, B.; Mitlin, D. Lithium ion battery applications of molybdenum disulfide (MoS2) nanocomposites. Energy Environ. Sci. 2014, 7, 209–231. [Google Scholar] [CrossRef]
  25. Wan, Z.M.; Shao, J.; Yun, J.J.; Zheng, H.Y.; Gao, T.; Shen, M.; Qu, Q.T.; Zheng, H.H. Core-Shell Structure of Hierarchical Quasi-Hollow MoS2 Microspheres Encapsulated Porous Carbon as Stable Anode for Li-Ion Batteries. Small 2014, 10, 4975–4981. [Google Scholar] [CrossRef]
  26. Li, J.Y.; Hou, Y.; Gao, X.F.; Guan, D.S.; Xie, Y.Y.; Chen, J.H.; Yuan, C. A three-dimensionally interconnected carbon nanotube/layered MoS2 nanohybrid network for lithium ion battery anode with superior rate capacity and long-cycle-life. Nano Energy 2015, 16, 10–18. [Google Scholar] [CrossRef]
  27. Guo, Y.; Li, S.; Fang, Q.Y.; Zuo, J.L.; Liu, M.; Zhang, J. An integrated electrode based on nanoflakes of MoS2 on carbon cloth for enhanced lithium storage. Rsc. Adv. 2020, 10, 9335–9340. [Google Scholar] [CrossRef]
  28. Hu, S.; Chen, W.; Uchaker, E.; Zhou, J.; Cao, G.Z. Mesoporous Carbon Nanofibers Embedded with MoS2 Nanocrystals for Extraordinary Li-Ion Storage. Chem. A Eur. J. 2015, 21, 18248–18257. [Google Scholar] [CrossRef]
  29. Wang, X.X.; Tian, J.H.; Cheng, X.; Na, R.; Wang, D.D.; Shan, Z.Q. Chitosan-Induced Synthesis of Hierarchical Flower Ridge-like MoS2/N-Doped Carbon Composites with Enhanced Lithium Storage. ACS Appl. Mater. Interfaces 2018, 10, 35953–35962. [Google Scholar] [CrossRef]
  30. Anwer, S.; Huang, Y.X.; Li, B.S.; Govindan, B.; Liao, K.; Cantwell, W.J.; Wu, F.; Chen, R.J.; Zheng, L.X. Nature-Inspired, Graphene-Wrapped 3D MoS2 Ultrathin Microflower Architecture as a High-Performance Anode Material for Sodium-Ion Batteries. ACS Appl. Mater. Interfaces 2019, 11, 22323–22331. [Google Scholar] [CrossRef]
  31. Zhang, X.E.; Chen, X.; Ren, H.J.; Diao, G.W.; Chen, M.; Chen, S.W. Bowl-like C@MoS2 Nanocomposites as Anode Materials for Lithium-Ion Batteries: Enhanced Stress Buffering and Charge/Mass Transfer. ACS Sustain. Chem. Eng. 2020, 8, 10065–10072. [Google Scholar] [CrossRef]
  32. Sari, F.N.I.; Ting, J.M. High performance asymmetric supercapacitor having novel 3D networked polypyrrole nanotube/N-doped graphene negative electrode and core-shelled MoO3/PPy supported MoS2 positive electrode. Electrochim. Acta 2019, 320, 11. [Google Scholar]
  33. Zhang, S.; Yu, X.; Yu, H.; Chen, Y.; Gao, P.; Li, C.; Zhu, C. Growth of Ultrathin MoS2 Nanosheets with Expanded Spacing of (002) Plane on Carbon Nanotubes for High-Performance Sodium-Ion Battery Anodes. ACS Appl. Mater. Interfaces 2014, 6, 21880–21885. [Google Scholar] [CrossRef]
  34. Zhao, B.C.; Zhou, J.F.; Si, J.G.; Fang, Z.T.; Li, K.Z.; Ma, H.Y.; Dai, J.M.; Zhu, X.B.; Sun, Y.P. Glucose-Induced Synthesis of 1T-MoS2/C Hybrid for High-Rate Lithium-Ion Batteries. Small 2019, 15, 1805420. [Google Scholar]
  35. Hu, S.; Chen, W.; Zhou, J.; Yin, F.; Uchaker, E.; Zhang, Q.F.; Cao, G.Z. Preparation of carbon coated MoS2 flower-like nanostructure with self-assembled nanosheets as high-performance lithium-ion battery anodes. J Mater. Chem. A 2013, 24, 125–130. [Google Scholar] [CrossRef]
  36. Bindumadhavan, K.; Srivastava, S.K.; Mahanty, S. MoS2-MWCNT hybrids as a superior anode in lithium-ion batteries. Chem. Commun. 2013, 49, 1823–1825. [Google Scholar] [CrossRef] [PubMed]
  37. Yu, H.; Ma, C.; Ge, B.; Chen, Y.; Xu, Z.; Zhu, C.; Li, C.; Ouyang, Q.; Gao, P.; Li, J.; et al. Three-Dimensional Hierarchical Architectures Constructed by Graphene/MoS2 Nanoflake Arrays and Their Rapid Charging/Discharging Properties as Lithium-Ion Battery Anodes. Chem. A Eur. J. 2013, 19, 5818–5823. [Google Scholar] [CrossRef] [PubMed]
  38. Chang, K.; Chen, W.; Ma, L.; Li, H.; Li, H.; Huang, F.; Xu, Z.; Zhang, Q.; Lee, J.-Y. Graphene-like MoS2/amorphous carbon composites with high capacity and excellent stability as anode materials for lithium ion batteries. J. Mater. Chem. 2011, 21, 6251–6257. [Google Scholar] [CrossRef]
  39. Zhang, X.; Zhao, R.F.; Wu, Q.H.; Li, W.L.; Shen, C.; Ni, L.B.; Yan, H.; Diao, G.W.; Chen, M. Petal-like MoS2 Nanosheets Space-Confined in Hollow Mesoporous Carbon Spheres for Enhanced Lithium Storage Performance. ACS Nano 2017, 11, 8429–8436. [Google Scholar] [CrossRef] [PubMed]
  40. Huang, G.C.; Chen, T.; Chen, W.X.; Wang, Z.; Chang, K.; Ma, L.; Huang, F.H.; Chen, D.Y.; Lee, J.Y. Graphene-Like MoS2/Graphene Composites: Cationic Surfactant-Assisted Hydrothermal Synthesis and Electrochemical Reversible Storage of Lithium. Small 2013, 9, 3693–3703. [Google Scholar] [CrossRef]
  41. Shao, J.; Qu, Q.; Wan, Z.; Gao, T.; Zuo, Z.; Zheng, H. From Dispersed Microspheres to Interconnected Nanospheres: Carbon-Sandwiched Monolayered MoS2 as High-Performance Anode of Li-Ion Batteries. ACS Appl. Mater. Interfaces 2015, 22927–22934. [Google Scholar] [CrossRef] [PubMed]
  42. 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] [PubMed]
  43. Hu, L.; Ren, Y.; Yang, H.; Xu, Q. Fabrication of 3D Hierarchical MoS2/Polyaniline and MoS2/C Architectures for Lithium-Ion Battery Applications. ACS Appl. Mater. Interfaces 2014, 6, 14644–14652. [Google Scholar] [CrossRef]
  44. McCreary, K.M.; Hanbicki, A.T.; Robinson, J.T.; Cobas, E.; Culbertson, J.C.; Friedman, A.L.; Jernigan, G.G.; Jonker, B.T. Large-Area Synthesis of Continuous and Uniform MoS2 Monolayer Films on Graphene. Adv. Funct. Mater. 2014, 24, 6449–6454. [Google Scholar] [CrossRef]
  45. Jie, X.; Wang, X.; Yang, X.Q.; Xun, S.; Gao, L.; Phillip, K.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]
  46. Wang, M.; Li, G.; Xu, H.; Qian, Y.; 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]
  47. He, G.; Evers, S.; Liang, X.; Cuisinier, M.; Garsuch, A.; Nazar, L.F. Tailoring porosity in carbon nanospheres for lithium-sulfur battery cathodes. ACS Nano 2013, 7, 10920–10930. [Google Scholar] [CrossRef] [PubMed]
  48. Zhu, Y.; Xu, Y.; Liu, Y.; Luo, C.; Wang, C. Comparison of electrochemical performances of olivine NaFePO4 in sodium-ion batteries and olivine LiFePO4 in lithium-ion batteries. Nanoscale 2013, 5, 780–787. [Google Scholar] [CrossRef]
  49. Cao, X.H.; Shi, Y.M.; Shi, W.H.; Rui, X.H.; Yan, Q.Y.; Kong, J.; Zhang, H. Preparation of MoS2-Coated Three-Dimensional Graphene Networks for High-Performance Anode Material in Lithium-Ion Batteries. Small 2013, 9, 3433–3438. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic illustration of the formation mechanism of the hierarchical MoS2/C quasi-hollow microspheres.
Figure 1. Schematic illustration of the formation mechanism of the hierarchical MoS2/C quasi-hollow microspheres.
Polymers 13 00837 g001
Figure 2. X-ray diffraction (XRD) patterns of the synthesized pristine MoS2 and MoS2/C composites prepared with different amounts of glucose (a.u. is arbitrary unit).
Figure 2. X-ray diffraction (XRD) patterns of the synthesized pristine MoS2 and MoS2/C composites prepared with different amounts of glucose (a.u. is arbitrary unit).
Polymers 13 00837 g002
Figure 3. Scanning electron microscopy (SEM) image of (a,b) pristine MoS2, (c,d) MoS2/C-1, (e,f) MoS2/C-2, and (g,h) MoS2/C-3 composites.
Figure 3. Scanning electron microscopy (SEM) image of (a,b) pristine MoS2, (c,d) MoS2/C-1, (e,f) MoS2/C-2, and (g,h) MoS2/C-3 composites.
Polymers 13 00837 g003
Figure 4. Transmission electron microscope (TEM) images of (ac) MoS2/C-2 microspheres at different magnifications; (d) electron diffraction pattern of MoS2/C-2.
Figure 4. Transmission electron microscope (TEM) images of (ac) MoS2/C-2 microspheres at different magnifications; (d) electron diffraction pattern of MoS2/C-2.
Polymers 13 00837 g004
Figure 5. (a) Mo 3d spectra and (b) S 2p X-ray photoelectron spectroscopy (XPS) spectra of the MoS2/C-2 composite.
Figure 5. (a) Mo 3d spectra and (b) S 2p X-ray photoelectron spectroscopy (XPS) spectra of the MoS2/C-2 composite.
Polymers 13 00837 g005
Figure 6. Cyclic voltammograms (CV) of (a) pristine MoS2, (b) MoS2/C-1, (c) MoS2/C-2, and (d) MoS2/C-3 composite electrodes; galvanostatic charge/discharge profiles of (e) pristine MoS2, (f) MoS2/C-1, (g) MoS2/C-2, and (h) MoS2/C-3 electrodes.
Figure 6. Cyclic voltammograms (CV) of (a) pristine MoS2, (b) MoS2/C-1, (c) MoS2/C-2, and (d) MoS2/C-3 composite electrodes; galvanostatic charge/discharge profiles of (e) pristine MoS2, (f) MoS2/C-1, (g) MoS2/C-2, and (h) MoS2/C-3 electrodes.
Polymers 13 00837 g006
Figure 7. (a) Cycling performance of the bulk MoS2 and MoS2/C electrodes measured in the voltage range of 0.01–3.0 V at a current density of 100 mA g−1; (b) rate capability of the bulk MoS2 and MoS2/C electrodes between 0.01 and 3 V at different current densities.
Figure 7. (a) Cycling performance of the bulk MoS2 and MoS2/C electrodes measured in the voltage range of 0.01–3.0 V at a current density of 100 mA g−1; (b) rate capability of the bulk MoS2 and MoS2/C electrodes between 0.01 and 3 V at different current densities.
Polymers 13 00837 g007
Figure 8. Electrochemical impedance spectra of the bulk MoS2 and MoS2/C electrodes before cycling (a) and after cycling 50 cycles (b) at a current density of 100 mA/g.
Figure 8. Electrochemical impedance spectra of the bulk MoS2 and MoS2/C electrodes before cycling (a) and after cycling 50 cycles (b) at a current density of 100 mA/g.
Polymers 13 00837 g008
Table 1. Summary of electrochemistry properties of MoS2/C flexible electrodes.
Table 1. Summary of electrochemistry properties of MoS2/C flexible electrodes.
MaterialsCurrent Density
mA g−1
Cyclic NumberSpecific Capacity
mAh g−1
References
C@MoS2 microsphere100100652[25]
Bowl-like C@ MoS2100100798[31]
Flower-like MoS2/C10050834[35]
MoS2-MWCNT 10030938[36]
MoS2/C nanoflowers10050888.1[43]
MoS2/C-2100200598This work
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, X.; Tan, J.; Li, X.; Zhang, C. Glucose-Assisted One-Pot Hydrothermal Synthesis of Hierarchical-Structured MoS2/C Quasi-Hollow Microspheres for High-Performance Lithium Ion Battery. Polymers 2021, 13, 837. https://doi.org/10.3390/polym13050837

AMA Style

Liu X, Tan J, Li X, Zhang C. Glucose-Assisted One-Pot Hydrothermal Synthesis of Hierarchical-Structured MoS2/C Quasi-Hollow Microspheres for High-Performance Lithium Ion Battery. Polymers. 2021; 13(5):837. https://doi.org/10.3390/polym13050837

Chicago/Turabian Style

Liu, Xingang, Jiang Tan, Xi Li, and Chuhong Zhang. 2021. "Glucose-Assisted One-Pot Hydrothermal Synthesis of Hierarchical-Structured MoS2/C Quasi-Hollow Microspheres for High-Performance Lithium Ion Battery" Polymers 13, no. 5: 837. https://doi.org/10.3390/polym13050837

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