Next Article in Journal
Visible Light-Induced Reactivity of Plasmonic Gold Nanoparticles Incorporated into TiO2 Matrix towards 2-Chloroethyl Ethyl Sulfide
Previous Article in Journal
Effect of Full Temperature Field Environment on Bonding Strength of Aluminum Alloy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Solid-State Synthesis of Layered MoS2 Nanosheets with Graphene for Sodium-Ion Batteries

Centre for Materials for Electronics Technology (C-MET), Ministry of Electronics and Information Technology (MeitY), Panchavati, Pune 411008, India
*
Author to whom correspondence should be addressed.
Crystals 2021, 11(6), 660; https://doi.org/10.3390/cryst11060660
Submission received: 10 March 2021 / Revised: 8 April 2021 / Accepted: 9 April 2021 / Published: 10 June 2021
(This article belongs to the Special Issue Synthesis, Properties and Applications of 2D Materials)

Abstract

:
Sodium-ion batteries have potential as energy-storage devices owing to an abundant source with low cost. However, most electrode materials still suffer from poor conductivity, sluggish kinetics, and huge volume variation. It is still challenging to explore apt electrode materials for sodium-ion battery applications to avoid the pulverization of electrodes induced by reversible intercalation of large sodium ions. Herein, we report a single-step facile, scalable, low-cost, and high-yield approach to prepare a hybrid material; i.e., MoS2 with graphene (MoS2-G). Due to the space-confined effect, thin-layered MoS2 nanosheets with a loose stacking feature are anchored with the graphene sheets. The semienclosed hybrid architecture of the electrode enhances the integrity and stability during the intercalation of Na+ ions. Particularly, during galvanostatic study the assembled Na-ion cell delivered a specific capacity of 420 mAhg−1 at 50 mAg−1, and 172 mAhg−1 at current density 200 mAg−1 after 200 cycles. The MoS2-G hybrid excels in performance due to residual oxygen groups in graphene, which improves the electronic conductivity and decreases the Na+ diffusion barrier during electrochemical reaction, in comparison with a pristine one.

1. Introduction

Nowadays, lithium-ion batteries (LIBs) play a vital role in energy-storage applications. However, lithium reserves are limited, which is affecting the economics of LIBs. Therefore, focus has been directed to development of new-generation energy-storage systems for increasing demands of portable electronic products and the automobile sector. Recently, sodium-ion batteries (SIBs) are attracting more attention due to their low cost and suitable redox potential (–2.71 V vs. SHE), as well as abundant resources of sodium [1,2,3]. Therefore, sodium-ion energy-storage devices inclusive of Na-ion hybrid supercapacitors and Na-ion batteries have currently become a research hotspot [4,5,6]. However, Na+ (1.02 Å) has a larger ionic radii compared with Li+ (0.76 Å) because limited selective host materials are available [7]. Therefore, in SIBs, sluggish chemical kinetic behaviors with huge volume variation causes pulverization, leading to rapid capacity-fading [8].
Graphite is an good candidate for LIBs; however, it cannot be used for SIBs because sodium ions cannot intercalate in graphite. Hence, it is an important task for a designing electrode materials to host the Na+ ions with the larger radii [9,10]. For eradicating the obstacles in Na-ion batteries, the alloying- and conversion-based anode materials have been studied. In recent years, layered metal dichalcogenides such as NbSe2, SnSe2, MoSe2, WS2, MoS2, VS2, and SnS/SnS2 have attracted great attention because of their two-dimensional (2D) structures with enlarged interlayer spacing, which provide alternatives for developing improved anodes for SIBs [11,12,13,14,15,16,17]. Among these 2D dichalcogenides, layered structured MoS2 has attracted great attention due to its high theoretical capacity (670 mA h g−1), which is double than that of graphite and comparatively cheaper [18,19]. The covalent bond between molybdenum and sulfur with weak van der Waals forces in MoS2 is promising, owing to its large interlaminar distance, which smoothens the intercalation of Na ions with environmental benignity [20]. The spacing between the adjacent layers of MoS2 is 0.62 nm, which is larger than in graphite (0.35 nm) [21]. Due to these structural features, smooth intercalation and deintercalation of Na+ between the MoS2 planes is possible. However, the practical application of MoS2 is still hindered because of restacking of MoS2 layers, which further decreases the availability of active sites for Na+ ions [22]. In addition, the poor electronic conductivity of MoS2 limits the electrode reactions, and intercalation of Na ions induces a large mechanical strain, leading to pulverization and further exfoliation of MoS2, and rapid capacity-fading occurs [6]. In order to overcome these issues, one potential method is the construction of an MoS2 hybrid nanostructure. Some researchers reported hybrid nanostructures of MoS2 with Nb2O5, FeCo, HfO2, MoO2, TiO2, Fe3O4, etc., and explored these for a sodium-ion battery [1,7,23,24,25,26]. Some approaches, such as downscaling the bulk MoS2 into few layers, expanding the interlayer space of MoS2, adaptation with advanced carbonaceous materials, etc., have been developed to solve the as-discussed challenges. Among these, modifications with carbonaceous matrices was the most studied [13,21,25,26,27]. Within carbonaceous materials, graphene has excellent electrical, mechanical, and thermal properties. Thus, many researchers have focused on assembling 2D graphene (G) sheets to support the expanded MoS2 to construct different architectures. This architecture was found to play a very important role in increasing the electronic conductivity of MoS2, enlarging the adsorption energy of Na+ on the surface of MoS2 layer and maintaining the high diffusion mobility of Na+ [13,28].
In this work, by taking advantage of MoS2 and graphene integration of adsorption–intercalation–conversion Na+ ion-storage mechanisms, we synthesized a nanostructured MoS2-G composite via a facile solid-state method. This 2D MoS2 material was adroitly encapsulated into the graphene sheets and the fabricated nanostructure MoS2-G composite, which was evaluated for SIBs. Herein, the layered MoS2 with graphene can significantly trap electrons, thus changing the electron state on the multigrain boundary. This improves the charge transmission and increases the capacity. When MoS2-G hybrids were explored as an anode material in SIBs, they demonstrated good rate capability with cycling stability.

2. Materials and Methods

2.1. Experimental

In a typical synthesis, the in situ solid state method was used to synthesize MoS2 and MoS2-graphene (MoS2-G) nanocomposites. The analytical grade thiourea and diammonium molybdate ((NH4)2MoO4) were taken in a ratio of 1:4 mol, and the graphene oxide was used as received. All precursors were thoroughly ground with a mortar and pestle, and the mixture was heated at 550 °C under an Ar atmosphere for 3 h. The graphene oxide content in the MoS2-G nanocomposites were 1, 2.5, 5, and 10 wt %, which were denoted as MS-1, MS-2, MS-3, and MS-4, respectively. Further, the pure MoS2 prepared without graphene oxide was denoted as MS-0.

2.2. Materials Characterization

The crystal structures and phases of the anode nanostructures were examined with the powder X-ray diffraction technique (XRD, Bruker Advanced D8 system, Karlsruhe, Germany) using a Cu Kα radiation source in a 2θ range from 20 to 80° at ambient temperature. The morphological and microstructural analysis of the as-synthesized nanostructures were conducted with field emission scanning electron microscopy (FESEM, Hitachi, S-4800, Kyoto, Japan) and field emission transmission electron microscopy (FETEM by JEOL; JEM-2200FS Kyoto, Japan). The surface chemical composition was studied with X-ray photoelectron spectroscopy (XPS, Thermo Fisher Scientific Co., Theta Probe, Waltham, MA, USA). Room-temperature micro-Raman scattering was performed using a HR 800-Raman.

2.3. Electrode Fabrication and Electrochemical Measurement

To perform electrochemical measurements for Na ions, 2032-type coin cells were fabricated. All the CR2032-type coin cells were assembled in a glove box (MTI, USA) filled with Ar gas. Metallic sodium foil and quartz filter paper were used as the counter electrode and separator, respectively. A ratio of 1M NaClO4 in ethylene carbonate (EC):diethyl carbonate (DEC) (1:1 in volume) with 5% fluoroethylene carbonate was used as an electrolyte. The electrodes were prepared by mixing active material (80 wt %), conducting carbon (10 wt %) and polyvinylidene fluoride (PVDF, 10 wt %) dissolved in N-methyl-2-pyrrolidone (NMP), subsequently coated on the copper foil, and dried in a vacuum oven at 120 °C for 12 h. After drying, the coated tapes were cut into circular discs with a diameter of 16 mm, and were further used as working electrodes. Cyclic voltammetry behavior of the half cells was tested on the Autolab potentiostat/galvanostat (Metrohm Autolab) in a voltage range of 0.01 V to 3 V. The galvanostatic charge–discharge behavior was tested on an MTI battery analyzer (vs. Na/Na+) at room temperature. Electrochemical impedance spectroscopy (EIS) was carried out using an amplitude of 5 mV with a frequency range of 0.1 Hz to 1 MHz.

3. Results

Figure 1 shows the XRD patterns of pure MoS2 (MS-0) and MoS2-G (MS2–4) that were recorded and compared to analyze phase purity. Peaks at 2θ values 13.44, 33.08, 39.47, 58.57, and 69.44 were assigned to the (002), (100), (103), (110), and (220) lattice planes of hexagonal 2H-MoS2, respectively (JCPDS No. 37-1492). The XRD graph demonstrates that the (002@14.37) peak had noticeably shifted toward lower angles (002@13.44°). The shift implied an increase of the interlayer distance, which helps during intercalation of Na ions. According to Bragg’s law, the(002) plane interlayer distance of the MoS2 sample was calculated and observed to be ~0.63 nm, which was larger than that of standard data (0.61 nm). The (002) plane corresponding to the peak at 13.44° suggested a well-stacked layered structure. The (002) plane was broad and weak, which revealed the existence of a few-layer structure of the MoS2 nanosheets [21,29,30,31]. The purity of the nanocomposite was confirmed via XRD.
Field emission scanning electron microscopy (FESEM) was used to investigate the size and morphology of the synthesized samples. Figure 2 illustrates the amassed layered MoS2 sheets with a loose stacking feature. During the synthesis, clusters of MoS2 sheets were formed that were a few nanometers in thickness. The MoS2 sheets could be thin MoS2 layers curled up by the temperature annealing. With the presence of graphene oxide in the reaction mixture, the MoS2-G composites exhibited a curved thin flaky appearance. This was an indication that the MoS2 layers were well supported on the curved graphene surface. From the FESEM, it was quite clear that at a higher concentration of graphene, the MoS2 sheets might be sandwiched in between graphene layers.
Further, a detailed structural study of the samples was conducted by field emission transmission electron microscopy (FETEM). Figure 3 shows the typical FETEM images of the MS-2 sample. These clearly demonstrate the formation of MoS2 with a sheetlike structure and ultrathin layers. The typical few-layered MoS2 nanosheets are clearly seen in Figure 2a,b. The individual nanosheets seemed to be transparent, which also indicated their thinness. These nanosized layered MoS2 sheets with a loose stacking feature were anchored with the graphene sheets. The high-resolution FETEM image (Figure 3c) shows distinct lattice fringes of the MS-2 sample with d-spacing of 0.63 nm, which corresponded to the (002) lattice plane of the layered MoS2 (JCPDS: 37-1492). Figure 3d shows the semicrystalline nature of the MoS2 along with the (100) and (106) planes.
The Raman spectra of MoS2 (MS-0) and MoS2-G (MS-2) are shown in Figure 4. The Raman peaks appear at 379 and 402 cm−1, corresponding to the E12g and A1g modes of the hexagonal MoS2 crystal, respectively. The E12g mode involves the in-layer displacement of Mo and S atoms, whereas the A1g mode involves the out-of-layer symmetric displacements of S atoms along the c axis [32,33]. Several researchers have found that single-layer MoS2 prepared by different methods would display an A1g Raman peak at 402–403 cm−1 [29,34,35]. Figure 4 is in excellent agreement with the characteristics of single-layer MoS2, i.e., an A1g peak at 402 cm−1, which confirms that the MoS2 sheets in the composite were single layered. Additionally, in the MS-2 sample, the D (disordered) band and the G (graphite) band of carbon at around 1360 and 1592 cm−1, respectively, belong to the graphene; while the Raman spectrum of pure MoS2 showed the characteristic peak of MoS2 without the D and G bands of carbon.
Chemical states and composition of the material was investigated using XPS. Figure 5a–c is a series of high-resolution XPS spectra of a typical sample of MS-2 for Mo, S, and C, respectively. The binding energy level of Mo 3D spectra at 229.4 and 232.6 eV depicts Mo 3d5/2 and 3d3/2, respectively. (Figure 5a) [36,37]. Figure 5b shows binding energies at 162.2 eV and 163.3 eV for S 2p3/2 and S 2p1/2, respectively [21]. These values confirmed the Mo in (IV+) oxidation state representing the presence of sulfur associated with Mo4+, which resonates with previous reports [38]. The peak of S2s centered at 226.5 eV was in the Mo 3D spectrum, and demonstrated +4 and −2 valences for Mo and S, respectively, which were maintained from S–Mo–S bonds. Figure 5c shows the C 1s spectra were deconvoluted into four separate Gaussian fitted peaks. The peak centered at 284.8 eV exhibited the conjugated sp2 C=C bonding in graphitic structure, while the other two peaks located at 284.5 eV and 286 eV were assigned to multifarious oxygen-containing functional groups such as C–O and C=O, respectively. The C1s spectrum shows bimodal distribution of peaks, which indicated significantly high oxygen contribution, which is beneficial for electrochemical reaction. The good performance of MoS2-G hybrids was due to graphene with residual oxygen-containing groups, which consequently improved the electronic conductivity of graphene and decreased the Na+ diffusion barrier during the MoS2-G interfaces, in comparison with the pristine one.
The MoS2-G layers were synthesized using diammonium molybdate, thiourea, and graphene oxide (GO). Usage of GO nanosheets as a substrate for the nucleation and subsequent growth of MoS2 have been reported [39]. Initially, H2S is formed via decomposition of thiourea, which is further decomposed into sulfur, which is attached on the surface of graphene through nucleation. Many researchers have reported the growth mechanism of graphene-based materials that used easy adsorption of cations on the GO surface, owing to the presence of negative charges on their surface [40,41]. During the calcination process, after dissociation of diammonium molybdate, molybdenum oxide and ammonia were formed. Furthermore, the molybdenum oxide reacted with sulfur, resulting in the formation of MoS2 nanosheets (2H phase) at 550 °C. During the reaction, reduction of graphene oxide (CnHn(OH)n) took place and released –H and =OH, which helped in the combustion of carbon and other combustible compounds.
( NH 4 ) 2 MoO 4 + 4 NH 2   CSNH 2 + GO MoS 2 + RGO + nH 2 O + 10 NH 3 + nCO 2 + 2 H 2 S
The electrochemical behaviors of the typical samples of MS-0 and MS-2 were first evaluated by cyclic voltammetry (CV) measurement. As displayed in Figure 6a,b, three reductive peaks of MoS2 at 0.1 mV were well observed in the first cycle. The peaks located at 0.74 V corresponded to the insertion of Na+ into MoS2, and the peaks located at around 1.3 V corresponded to the formation of Mo and Na2S. The anodic peak located at 1.81 V was attributed to oxidation reaction from Mo to MoS2. In the subsequent cycles, the anodic curves showed characteristics similar to those of the first cycle, whereas the cathodic curves showed a difference. This change arose due to the formation of SEI film during the first cycle. The CV curves at the third cycle were stable and similar to those of the second cycle, which demonstrated reversibility and a stable sodiation/desodiation process of the MoS2, as per the following equations.
At counter electrode:
N a x N a + x e
During discharge (intercalation at working electrode):
x N a + + x e + M o S 2 N a x M o S 2
During charge (conversion reaction at working electrode):
N a x M o S 2 + 4 N a + + 4 e M o + 2 N a 2 S + x N a
where x represents the number of moles of corresponding intercalating Na ions or electrons in the active material.
Figure 6c shows the AC impedance spectra of the MS-0 and MS-2 electrodes, which were measured at the identical condition in the frequency range of 0.01 to 100 kHz at OCV. The charge-transfer resistance (Rct) determined in the medium-frequency region from the semicircle was 750 and 70 Ω for MS-0 and MS-2, respectively. The MS-2 sample had a lower charge-transfer resistance due to the incorporation of graphene, which increased the electronic conductivity of MoS2.
The exchange-current density ( i 0) is inversely proportional to the charge transfer resistance. These charge-transfer resistance results revealed that the MS-2 sample had a higher exchange-current density than MS-0, which confirmed that graphene effectively increased the Na+ ion diffusion and electronic conductivity, which controlled the interfacial resistance between particles during the electrochemical reactions.
Additionally, to investigate the effect of graphene on MoS2 in terms of Na+ ion diffusion calculated from the relationship between Zre and ω–1/2 in the low-frequency region, we used the following equations:
Z r e = R s + R c t + σ w ω 1 / 2
D = R 2 T 2 2 A 2 n 4 F 4 C 2 σ w 2
where ω is angular frequency in the low frequency region, σw represents the Warburg impedance coefficient, D is the Na+ diffusion coefficient, R is the gas constant, T is the absolute temperature, A is the area of electrode surface, n is the number of the electrons per molecule participating in the electronic transfer reaction, F is the Faraday constant, and C is the molar concentration of Na+. Figure 6d shows the correlation between Zre and ω−1/2 for the MS-0 and MS-2 samples in the low-frequency region; the slope of the fitted line is the Warburg coefficient σ. The diffusion coefficients of sodium ions were calculated to be 1.61 × 10−13 and 1.84 × 10−12 for MS-0 and MS-2, respectively. The results clearly showed that due to graphene, the Warburg coefficient decreased, but the Na ion diffusion coefficient increased. As compared to pristine, the MoS2-G enhanced the electronic conductivity, as well as sodium-ion diffusion, and lowered charge-transfer resistance, which was constructive for the electrochemical performance of the electrodes. The electronic conductivity of doped MS-2 was observed to be higher; therefore, electrochemical polarization decreased as compared to pristine.
Figure 7a,b demonstrate the charge–discharge profiles of the MoS2-G composite at a current rate of 50 mA g−1, in which three plateaus at around 0.7 V, 0.6 V, and 0.1 V in the first discharge process, and one plateau at 1.8 V in the first charge process, are clearly observed. The second and third cycles showed similar characteristics, which was in good accordance with the CV results. The first discharge capacities of samples MS-0 and MS-2 were observed at 807.57 and 641.45 mAhg−1, with reversible charging capacities of 305.21 and 420.43 mAhg−1; the respective retention efficiencies were observed to be 37 and 65.52%. The lower retention for the MS-0 samples in the first cycle was attributed to the formation of more solid electrolyte interface (SEI) film compared to MS-2. The SEI layer formation took place due to the decomposition of electrolyte, as well as irreversible electrochemical reactions of MoS2.
To check the rate capability of the MoS2 (MS-0 and MS-2), the detailed galvanostatic discharge/charge behaviors were examined at different current densities; i.e., 50, 100, 250, 500, and 1000 mAg−1, and retained to 50 mAg−1 for the last five cycles, as shown in Figure 8a. Pure MoS2 (MS-0) exhibited 305 reversible capacity at 50 mAg−1, and it was much less at 1000 mAg−1. However, within different percentages of graphene, MS-2 exhibited better reversible capacity; i.e., 420 and 78 mAhg−1 at 50 and 1000 mAg−1, respectively. With increased current density, the specific capacities continuously dropped for pure samples, while the MS-2 sample showed good rate performance. When the current reverted back to 50 mAg−1, the MS-0 electrode delivered 121.99 mAhg−1, while the graphene-modified MoS2 (MS-2) exhibited 316.14 mAhg−1. These results demonstrated that the cycling and rate performance of MoS2 was greatly improved by optimum graphene doping. This revealed that the MS-2 nanosheets with 2.5% graphene could sustain various current rates while keeping their stable structure. An optimum amount of electronic conductor; i.e., graphene, gave stability to the active material, and also provided three-dimensional electronic channels that were favorable for the diffusion during the discharging/charging process. However, a higher percentage of graphene limited the active sites of MoS2 for intercalation of Na ions due to the shielding effect of graphene layers, which might be responsible for the lower capacity. From the FESEM, it was quite clear that at higher concentrations of graphene, the MoS2 sheets might be sandwiched in between graphene layers. Ultimately, it decreased ionic mobility prior to electronic conductivity. Hence, this experimental evidence clearly showed the necessity of optimum graphene.
The enhanced performance of MS-2 (MoS2-G) nanosheets was due to the expanded d-spacing of MoS2 layers, which facilitated smooth intercalation of Na+ ions. Simultaneously, nanosheets with a few-layer structure were prevented from volume expansion and pulverization of the electrode. Due to the space-confined effect, nanosized layered MoS2 sheets with a loose stacking feature were anchored with the graphene sheets. Due to the mechanical and electrical properties of graphene, the electron transport accelerated, resulting in fast kinetics within the MoS2-G composites. The semienclosed hybrid architecture enhanced the stability and integrity of the electrode structure during intercalation of Na+ ions.
Figure 8b shows the cycling performance of pure MoS2 (MS-0) and MoS2-G (MS-2, considering better capacity), to check the stability of electrodes at 200 mAg−1 in the voltage range of 0.01–3 V. During this study, the coin cell was activated at current density 50 mAg−1, and further, the electrode was cycled at 200 mAg−1. The specific capacity of the pure MoS2 anode at 100 mAg−1 was 98.79 mAhg−1, and in the subsequent cycles, its capacity dropped continuously, whereas the MoS2-G (MS-2) electrode delivered a specific capacity of 172 mA h g−1 after 200 cycles, with a capacity retention of 70%. The coulombic efficiencies of respective cycles were maintained at around 99% in the subsequent cycles, with marginal degradation.
On the basis of the comparative results for the electrochemical performance of the as-prepared materials, the good capacity, good rate behavior, and cycling stability of MoS2-G were attributed to the small size and ultrathin thickness of the MoS2 nanosheets together with the conductive properties of graphene. The expanded interlayer space of MoS2 facilitated the smoother intercalation of large-size Na ions, and also eased the volume expansion and pulverization of MoS2 during Na ion insertion. Optimum graphene provided a continuous pathway for electron transport and facilitated the storage and diffusion of Na ions. The 2D nanosheets of MoS2 with graphene could effectively buffer the stress induced by the large volume variation of electrode materials during the sodiation/desodiation process, and thereby benefited the structural stability and integrity of the electrode.

4. Conclusions

In summary, ultrathin MoS2-graphene nanosheets were fabricated by a facile solid-state method and checked for feasibility in Na-ion batteries. Due to its structure, the electrode achieved a good reversible specific capacity of 420 mAhg−1 at 50 m·Ag−1, and exhibited 78 mAhg−1 at 1000 mAg−1. The improved performance of MoS2-G over pristine MoS2 was attributed to the expanded d-spacing, which facilitated smooth intercalation of Na+ ions. The nanosized layered MoS2 sheets with a loose stacking feature, which were anchored on the graphene sheets, accelerated the ion and electron transport, thus resulting in fast kinetics for the electrochemical reactions. This hybrid structure prevented volume expansion and pulverization of the electrode during repeated intercalation of Na+ ions. The semienclosed hybrid architecture enhanced the stability and integrity of the electrode structure during intercalation of Na+ ions.

Author Contributions

Methodology, writing—original draft preparation, U.C.; formal analysis, C.U., M.K.; conceptualization, supervision, validation, B.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Ministry of Science and Technology (No. SR/WOS-A/CS-73/2017).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article.

Acknowledgments

The authors would like to thank the Ministry of Science and Technology (No. SR/WOS-A/CS-73/2017) for financial support. The authors also would like to thank C-MET, Pune, Ministry of Electronics and Information Technology (MeitY), and the Government of India for providing research facilities, and the Nanocrystalline Materials Group for kind support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Deng, Q.; Chen, F.; Liu, S.; Bayaguud, A.; Feng, Y.; Zhang, Z.; Fu, Y.; Yu, Y.; Zhu, C. Advantageous Functional Integration of Adsorption-Intercalation-Conversion Hybrid. Mechanisms in 3D Flexible Nb2O5@Hard Carbon@MoS2@Soft Carbon Fiber Paper Anodes for Ultrafast and Super-Stable Sodium Storage. Adv. Funct. Mater. 2020, 30, 1908665. [Google Scholar]
  2. Kim, H.; Kim, H.; Ding, Z.; Lee, M.H.; Lim, K.; Yoon, G.; Kang, K. Recent Progress in Electrode Materials for Sodium-Ion Batteries. Adv. Energy Mater. 2016, 6, 1600943. [Google Scholar] [CrossRef] [Green Version]
  3. Choi, J.W.; Aurbach, D. Promise and reality of post-lithium-ion batteries with high energy densities. Nat. Rev. Mater. 2016, 1, 16013. [Google Scholar] [CrossRef]
  4. Zeng, L.; Luo, F.; Chen, X.; Xu, L.; Xiong, P.; Feng, X.; Luo, Y.; Chen, Q.; Wei, M.; Qian, Q. An ultra-small few-layer MoS2-hierarchical porous carbon fiber composite obtained via nanocasting synthesis for sodium-ion battery anodes with excellent long-term cycling performance. Dalton Trans. 2019, 48, 4149–4156. [Google Scholar] [CrossRef]
  5. Yang, R.; Sun, L.; Liu, W.; Zhang, Y.; Cui, Y.; Du, Y.; Liu, S.; Wang, H.; Huang, M. Bio-derived 3D TiO2 hollow spheres with a mesocrystal nanostructure to achieve improved electrochemical performance of Na-ion batteries in ether-based electrolytes. J. Mater. Chem. A 2019, 7, 3399–3407. [Google Scholar] [CrossRef]
  6. Zhang, Y.; Liu, W.; Wang, T.; Du, Y.; Cui, Y.; Liu, S.; Wang, H.; Liu, S.; Chen, M.; Zhou, J. Space-Confined Fabrication of MoS2@Carbon Tubes with Semienclosed Architecture Achieving Superior Cycling Capability for Sodium Ion. Storage Adv. Mater. Interfaces 2020, 7, 2000953. [Google Scholar] [CrossRef]
  7. Wang, C.; Yang, Q.; Qin, G.; Xiao, Y.; Duan, J. Unveiling a bimetallic FeCo-coupled MoS2 composite for enhanced energy storage. Nanoscale 2020, 12, 10532–10542. [Google Scholar] [CrossRef]
  8. Wei, X.; Zhang, Y.; Zhang, B.; Lin, Z.; Wang, X.; Hu, P.; Li, S.; Tan, X.; Cai, X.; Yang, W.; et al. Yolk-shell-structured zinc-cobalt binary metal sulfide @ N-doped carbon for enhanced lithium-ion storage. Nano Energy 2019, 64, 103899. [Google Scholar] [CrossRef]
  9. Li, Z.; Zhan, X.; Qi, S. A facile alkali metal hydroxide-assisted controlled and targeted synthesis of 1T MoS2 single-crystal nanosheets for lithium ion battery anodes. Nanoscale 2019, 11, 14857–14862. [Google Scholar] [CrossRef] [PubMed]
  10. Han, M.; Lin, Z.; Yu, J. Ultrathin MoS2 nanosheets homogenously embedded in a N,O-codoped carbon matrix for high-performance lithium and sodium storage. J. Mater. Chem. A 2019, 7, 4804–4812. [Google Scholar] [CrossRef]
  11. Subramanian, Y.; Veerasubramani, G.K.; Park, M.S.; Kim, D.W. Investigation of Layer Structured NbSe2 as an Intercalation Anode Material for Sodium-Ion. Hybrid. Capacitors. J. Electrochem. Soc. 2019, 166, A598–A604. [Google Scholar] [CrossRef]
  12. Su, D.; Dou, S.; Wang, G. WS2@graphene nanocomposites as anode materials for Na-ion batteries with enhanced electrochemical performances. Chem. Commun. 2014, 50, 4192–4195. [Google Scholar] [CrossRef]
  13. Sun, D.; Ye, D.; Liu, P.; Tang, Y.; Guo, J.; Wang, L.; Wang, H. MoS2/Graphene Nanosheets from Commercial Bulky MoS2 and Graphite as Anode Materials for High. Rate Sodium-Ion. Batteries. Adv. Energy Mater. 2018, 8, 1702383. [Google Scholar]
  14. Wu, D.; Wang, C.; Wu, M.; Chao, Y.; He, P.; Ma, J. Porous bowl-shaped VS2 nanosheets/graphene composite for high-rate lithium-ion storage. J. Energy Chem. 2020, 43, 24–32. [Google Scholar] [CrossRef] [Green Version]
  15. Zhou, T.; Pang, W.K.; Zhang, C.; Yang, J.; Chen, Z.; Liu, H.K.; Guo, Z. Enhanced Sodium-Ion. Battery Performance by Structural Phase Transition from Two-Dimensional Hexagonal-SnS2 to Orthorhombic-SnS. ACS Nano 2014, 8, 8323–8333. [Google Scholar] [PubMed]
  16. Zhang, F. SnSe2 Two Dimensional Anodes for Advanced Sodium Ion Batteries. Master’s Thesis, King Abdullah University of Science and Technology, Thuwal, Saudi Arabia, 2017. [Google Scholar]
  17. Zhang, Y.; Liu, Z.; Zhao, H.; Du, Y. MoSe2 nanosheets grown on carbon cloth with superior electrochemical performance as flexible electrode for sodium ion batteries. RSC Adv. 2016, 6, 1440–1444. [Google Scholar] [CrossRef]
  18. Wang, J.; Luo, C.; Gao, T.; Langrock, A.; Mignerey, A.C.; Wang, C. An Advanced MoS2/Carbon Anode for High-Performance Sodium-Ion Batteries. Small 2015, 11, 473–481. [Google Scholar] [CrossRef]
  19. Cao, X.; Shi, Y.; Shi, W.; Rui, X.; Yan, Q.; 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]
  20. Kumar, P.R.; Jung, Y.H.; Kim, D.K. High performance of MoS2 microflowers with a water-based binder as an anode for Na-ion batteries. RSC Adv. 2015, 5, 79845–79851. [Google Scholar]
  21. Zhang, P.; Qin, F.; Zou, L.; Wang, M.; Zhang, K.; Lai, Y.; Li, J. Few-layered MoS2/C with expanding d-spacing as a high-performance anode for sodium-ion batteries. Nanoscale 2017, 9, 12189–12195. [Google Scholar] [CrossRef]
  22. Park, S.K.; Lee, J.; Bong, S.; Jang, B.; Seong, K.D.; Piao, Y. Scalable Synthesis of Few-Layer MoS2 Incorporated into Hierarchical Porous Carbon Nanosheets for High.-Performance Li- and Na-Ion. Battery Anodes. ACS Appl. Mater. Interfaces 2016, 8, 19456–19465. [Google Scholar] [CrossRef]
  23. Ahmed, B.; Anjum, D.H.; Hedhili, M.N.; Alshareef, H.N. Mechanistic Insight into the Stability of HfO2-Coated MoS2 Nanosheet Anodes for Sodium Ion. Batteries. Small 2015, 11, 4341–4350. [Google Scholar] [CrossRef] [PubMed]
  24. Xiao, D.; Zhang, J.; Li, X.; Zhao, D.; Huang, H.; Huang, J.; Cao, D.; Li, Z.; Niu, C. Nanocarved MoS2–MoO2 Hybrids Fabricated Using in Situ Grown MoS2 as Nanomasks. ACS Nano 2016, 10, 9509–9515. [Google Scholar] [CrossRef] [PubMed]
  25. Ren, W.; Zhou, W.; Zhang, H.; Cheng, C. ALD TiO2-Coated Flower-like MoS2 Nanosheets on Carbon Cloth. as Sodium Ion. Battery Anode with Enhanced Cycling Stability and Rate Capability. ACS Appl. Mater. Interfaces 2017, 9, 487–495. [Google Scholar] [PubMed]
  26. Kong, D.; Cheng, C.; Wang, Y.; Huang, Z.; Liu, B.; Von Lim, Y.; Ge, Q.; Yang, H.Y. Fe3O4 quantum dot decorated MoS2 nanosheet arrays on graphite paper as free-standing sodium-ion battery anodes. J. Mater. Chem. A 2017, 5, 9122–9131. [Google Scholar] [CrossRef]
  27. Zhu, C.; Mu, X.; van Aken, P.A.; Yu, Y.; Maier, J. Single-layered ultrasmall nanoplates of MoS2 embedded in carbon nanofibers with excellent electrochemical performance for lithium and sodium storage. Angew. Chem. Int. Ed. Engl. 2014, 53, 2152–2156. [Google Scholar] [CrossRef]
  28. Xie, X.; Ao, Z.; Su, D.; Zhang, J.; Wang, G. MoS2/Graphene Composite Anodes with Enhanced Performance for Sodium-Ion. Batteries: The Role of the Two-Dimensional Heterointerface. Adv. Funct. Mater. 2015, 25, 1393–1403. [Google Scholar]
  29. Wang, Z.; Chen, T.; Chen, W.; Chang, K.; Ma, L.; Huang, G.; Chen, D.; Lee, J.Y. CTAB-assisted synthesis of single-layer MoS2–graphene composites as anode materials of Li-ion batteries. J. Mater. Chem. A 2013, 1, 2202–2210. [Google Scholar] [CrossRef]
  30. 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]
  31. Ramakrishna Matte, H.S.; Gomathi, A.; Manna, A.K.; Late, D.J.; Datta, R.; Pati, S.K.; Rao, C.N. MoS2 and WS2 Analogues of Graphene. Angew. Chem. Int. Ed. 2010, 49, 4059–4062. [Google Scholar] [CrossRef] [PubMed]
  32. Frey, G.L.; Tenne, R.; Matthews, M.J.; Dresselhaus, M.S.; Dresselhaus, G. Raman and resonance Raman investigation of MoS2 nanoparticles. Phys. Rev. B 1999, 60, 2883–2892. [Google Scholar] [CrossRef]
  33. Wang, Y.; Qu, Q.; Li, G.; Gao, T.; Qian, F.; Shao, J.; Liu, W.; Shi, Q.; Zheng, H. 3D Interconnected and Multiwalled Carbon@MoS2@Carbon Hollow Nanocables as Outstanding Anodes for Na-Ion. Batteries. Small 2016, 12, 6033–6041. [Google Scholar] [CrossRef] [PubMed]
  34. Lee, H.S.; Min, S.W.; Chang, Y.G.; Park, M.K.; Nam, T.; Kim, H.; Kim, J.H.; Ryu, S.; Im, S. MoS2 Nanosheet Phototransistors with Thickness-Modulated Optical Energy Gap. Nano Lett. 2012, 12, 3695–3700. [Google Scholar] [CrossRef]
  35. Lee, C.; Yan, H.; Brus, L.E.; Heinz, T.F.; Hone, J.; Ryu, S. Anomalous Lattice Vibrations of Single- and Few-Layer MoS2. ACS Nano 2010, 4, 2695–2700. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Chang, Y.H.; Wu, F.Y.; Chen, T.Y.; Hsu, C.L.; Chen, C.H.; Wiryo, F.; Wei, K.H.; Chiang, C.Y.; Li, L.J. Three-Dimensional Molybdenum Sulfide Sponges for Electrocatalytic Water Splitting. Small 2014, 10, 895–900. [Google Scholar] [CrossRef]
  37. Chen, X.; Zhang, Z.; Li, X.; Shi, C.; Li, X. Selective synthesis of metastable MoO2 nanocrystallites through a solution-phase approach. Chem. Phys. Lett. 2006, 418, 105–108. [Google Scholar] [CrossRef]
  38. Kawade, U.V.; Ambalkar, A.A.; Panmand, R.P.; Kalubarme, R.S.; Kadam, S.R.; Naik, S.D.; Kulkarni, M.V.; Kale, B.B. Silicon nanoparticle-sandwiched ultrathin MoS2–graphene layers as an anode material for Li-ion batteries. Mater. Chem. Front. 2019, 3, 587–596. [Google Scholar] [CrossRef]
  39. Li, Y.; Wang, H.; Xie, L.; Liang, Y.; Hong, G.; Dai, H. MoS2 Nanoparticles Grown on Graphene: An. Advanced Catalyst for the Hydrogen Evolution Reaction. J. Am. Chem. Soc. 2011, 133, 7296–7299. [Google Scholar]
  40. Chang, K.; Geng, D.; Li, X.; Yang, J.; Tang, Y.; Cai, M.; Li, R.; Sun, X. Ultrathin MoS2/Nitrogen-Doped Graphene Nanosheets with Highly Reversible Lithium Storage. Adv. Energy Mater. 2013, 3, 839–844. [Google Scholar] [CrossRef]
  41. Sahoo, N.G.; Pan, Y.; Li, L.; Chan, S.H. Graphene-Based Materials for Energy Conversion. Adv. Mater. 2012, 24, 4203–4210. [Google Scholar] [CrossRef]
Figure 1. XRD of pure MoS2 (MS-0) and MoS2-G (MS-1, MS-2, MS-3, and MS-4).
Figure 1. XRD of pure MoS2 (MS-0) and MoS2-G (MS-1, MS-2, MS-3, and MS-4).
Crystals 11 00660 g001
Figure 2. FESEM of (a) pure MoS2 (MS-0), (b) MoS2-G, (c,d) MS-1, (e,f) MS-2, (g,h) MS-3, and (i,j) MS-4 at different magnifications.
Figure 2. FESEM of (a) pure MoS2 (MS-0), (b) MoS2-G, (c,d) MS-1, (e,f) MS-2, (g,h) MS-3, and (i,j) MS-4 at different magnifications.
Crystals 11 00660 g002
Figure 3. (a,b) FETEM of MoS2 (MS-2) at different magnifications; (c) FETEM of the individual MoS2 layer; (d) corresponding selected area electron diffraction SAED patterns.
Figure 3. (a,b) FETEM of MoS2 (MS-2) at different magnifications; (c) FETEM of the individual MoS2 layer; (d) corresponding selected area electron diffraction SAED patterns.
Crystals 11 00660 g003
Figure 4. Raman spectra of pure MoS2 (MS-0) and MoS2-G (MS-2).
Figure 4. Raman spectra of pure MoS2 (MS-0) and MoS2-G (MS-2).
Crystals 11 00660 g004
Figure 5. XPS spectra for MoS2-G (MS-2): (a) Mo, (b) S, and (c) C.
Figure 5. XPS spectra for MoS2-G (MS-2): (a) Mo, (b) S, and (c) C.
Crystals 11 00660 g005
Figure 6. The first three consecutive CV curves of (a) MS-0 and (b) MS-2 at a scan rate of 0.1 mVs−1 (c) Electrochemical impedance spectrum at OCV (d) The relationship between Zre and ω−1/2.
Figure 6. The first three consecutive CV curves of (a) MS-0 and (b) MS-2 at a scan rate of 0.1 mVs−1 (c) Electrochemical impedance spectrum at OCV (d) The relationship between Zre and ω−1/2.
Crystals 11 00660 g006
Figure 7. Electrochemical properties: the initial charge–discharge profiles at 50 mAg−1 current density for (a) MS-0; (b) MS-2.
Figure 7. Electrochemical properties: the initial charge–discharge profiles at 50 mAg−1 current density for (a) MS-0; (b) MS-2.
Crystals 11 00660 g007
Figure 8. Electrochemical properties: (a) the rate performance of all samples at different current densities; and (b) the cycling performance of MS-0 and MS-2 between 0.01 and 3 V at a current rate of 200 mAg−1.
Figure 8. Electrochemical properties: (a) the rate performance of all samples at different current densities; and (b) the cycling performance of MS-0 and MS-2 between 0.01 and 3 V at a current rate of 200 mAg−1.
Crystals 11 00660 g008
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chothe, U.; Ugale, C.; Kulkarni, M.; Kale, B. Solid-State Synthesis of Layered MoS2 Nanosheets with Graphene for Sodium-Ion Batteries. Crystals 2021, 11, 660. https://doi.org/10.3390/cryst11060660

AMA Style

Chothe U, Ugale C, Kulkarni M, Kale B. Solid-State Synthesis of Layered MoS2 Nanosheets with Graphene for Sodium-Ion Batteries. Crystals. 2021; 11(6):660. https://doi.org/10.3390/cryst11060660

Chicago/Turabian Style

Chothe, Ujjwala, Chitra Ugale, Milind Kulkarni, and Bharat Kale. 2021. "Solid-State Synthesis of Layered MoS2 Nanosheets with Graphene for Sodium-Ion Batteries" Crystals 11, no. 6: 660. https://doi.org/10.3390/cryst11060660

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