Next Article in Journal
Strain Effects on the Electronic and Optical Properties of Kesterite Cu2ZnGeX4 (X = S, Se): First-Principles Study
Next Article in Special Issue
Nitrogen-Doped Reduced Graphene Oxide Supported Pd4.7Ru Nanoparticles Electrocatalyst for Oxygen Reduction Reaction
Previous Article in Journal
Mesoporous Zn/MgO Hexagonal Nano-Plates as a Catalyst for Camelina Oil Biodiesel Synthesis
Previous Article in Special Issue
Ammonium Ion Enhanced V2O5-WO3/TiO2 Catalysts for Selective Catalytic Reduction with Ammonia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Reduced Graphene-Oxide-Encapsulated MoS2/Carbon Nanofiber Composite Electrode for High-Performance Na-Ion Batteries

1
Department of Materials Science and Engineering, Korea Advanced Institute of Science and Technology (KAIST), Daejeon 34141, Korea
2
Department of Materials Science and Engineering, Chungnam National University, Daejeon 34134, Korea
3
School of Materials Science and Engineering, University of Ulsan, Ulsan 44776, Korea
*
Authors to whom correspondence should be addressed.
S.-H.C. and J.-H.K. contributed equally to this work.
Nanomaterials 2021, 11(10), 2691; https://doi.org/10.3390/nano11102691
Submission received: 15 September 2021 / Revised: 7 October 2021 / Accepted: 11 October 2021 / Published: 13 October 2021
(This article belongs to the Special Issue Nanomaterials for Energy Conversion and Catalytic Applications)

Abstract

:
Sodium-ion batteries (SIBs) have been increasingly studied due to sodium (Na) being an inexpensive ionic resource (Na) and their battery chemistry being similar to that of current lithium-ion batteries (LIBs). However, SIBs have faced substantial challenges in developing high-performance anode materials that can reversibly store Na+ in the host structure. To address these challenges, molybdenum sulfide (MoS2)-based active materials have been considered as promising anodes, owing to the two-dimensional layered structure of MoS2 for stably (de)inserting Na+. Nevertheless, intrinsic issues of MoS2—such as low electronic conductivity and the loss of active S elements after a conversion reaction—have limited the viability of MoS2 in practical SIBs. Here, we report MoS2 embedded in carbon nanofibers encapsulated with a reduced graphene oxide (MoS2@CNFs@rGO) composite for SIB anodes. The MoS2@CNFs@rGO delivered a high capacity of 345.8 mAh g−1 at a current density of 100 mA g−1 for 90 cycles. The CNFs and rGO were synergistically taken into account for providing rapid pathways for electrons and preventing the dissolution of S sources during repetitive conversion reactions. This work offers a new point of view to realize MoS2-based anode materials in practical SIBs.

1. Introduction

A battery of things (BoT)—first mentioned by Tony Seba, the author of “Clean Disruption of Energy and Transportation”—has come for modern society, requiring battery-powered devices everywhere [1]. Current lithium-ion batteries (LIBs), with a high energy density of ~250 Wh kg−1, have been considered to be promising energy storage systems combined with renewable energy technologies [2]. Despite the viability of LIBs, utilization of the state-of-the-art LIBs for an energy storage system (ESS) has been restricted due to finite reserves of Li+ on earth, which would increase the price of Li sources used for large-scale applications [3]. Among alternatives to LIBs, sodium-ion batteries (SIBs) with a similar battery chemistry (redox potential of −2.71 V vs. SHE) as LIBs have constituted an exciting avenue for advancing ESSs [4,5].
However, anode materials with low energy density have impeded the development of SIB technologies. To address this issue, many promising candidates (e.g., hard carbon, metal oxides/sulfides/selenides, etc.) with high capacities have been suggested so far [6,7,8,9,10,11,12]. However, carbonaceous materials show limited use due to their low specific capacity and less reversible capacity, and transition metal oxides (TMOs) need to be changed due to their low conductivity. As high-capacity anodes, sulfide-based materials such as FeS2 and Sn2S have been developed because transition metal sulfides (TMSs) have a higher electronic conductivity as well as an excellent ability to store Na+ [11,12]. Among them, molybdenum sulfide (MoS2)-based anode materials possessing large interlayer spaces (0.615 nm) between two-dimensional (2D) MoS2 slabs could allow facile (de)insertion of Na+ with few structural changes during conversion reactions (MoS2 + 4Na+ + 4e → Mo(metallic) + 2Na2S, a theoretical capacity of 670 mAh g−1), leading to great electrochemical performance [13,14]. However, MoS2 has an intrinsically low electronic conductivity, lowering the efficiency of (dis) charging SIB cells [15,16]. In addition, long channels of 2D MoS2 interlayers are the bottleneck to Na+ diffusion, degrading SIB performance toward Na+ storage [17,18]. Furthermore, the dissolution of sulfur atoms contributes to a large loss in the overall mass after the conversion reactions, which is detrimental to achieving excellent battery performance. Therefore, we need to find a strategy to overcome the problems of MoS2 mentioned above [19,20,21].
In this work, we successfully fabricated interlayer-enlarged MoS2 nanoflakes, which were doubly covered with carbon nanofibers (CNFs), and reduced graphene oxide (rGO) (MoS2@CNFs@rGO). The MoS2 nanoflakes were first confined in the CNFs by thermolysis, subsequently encapsulated by the rGO via electrostatic interaction and reduction processes. The interlayers of MoS2 nanoflakes in the CNFs were expanded during the thermolysis at 800 °C under the H2 atmosphere. Moreover, the synergy of CNFs and rGO not only increase the electronic conductivity of the composite but also prevent the loss of S, enabling the SIB cells to be operated reversibly with a high capacity of 345.8 mA g−1 for 90 cycles at a high current density of 100 mA g−1. With the support of experimental and analytical studies, the underlying reaction mechanism of MoS2@CNFs@rGO was investigated and proposed.

2. Materials and Methods

2.1. Experimental

Firstly, the MoS2@CNFs were synthesized by electrospinning and a thermolysis process [17]. For the electrospinning, a solution containing 15 wt % of ammonium tetrathiomolybdate ((NH4)2MoS4, ATTM, Alfa Aesar, Ward Hill, MA, USA) and 15 wt % of poly(styrene-acrylonitrile, SAN, Mw = 1,300,000) dissolved in 10 mL of N,N-dimethylformamide (DMF, Sigma-Aldrich, Burlington, MA, USA), was prepared on the hot plate by stirring the solution at 70 °C for 12 h. The solution was electrospun by applying a high voltage of 15 kV using an electrospinning machine (NanoNC, Seoul, South Korea). After the electrospinning, the electrospun NFs were thermally treated under H2/Ar (4/96, v/v) surroundings at 450 °C for 2 h and under a pure Ar (99.999%) atmosphere at 800 °C for 6 h, respectively [17,22]. These processes were performed to make the MoS2@CNFs. For wrapping the entire surfaces of MoS2@rGO with the rGO, the Ti-O-Ti-O atomic layers (sub-nm) were coated on the MoS2@rGO by using atomic layer deposition (ALD); this forms hydroxyl groups (OH-) of the atomic layers on MoS2@rGO for rGO wrapping [7,23]. Then, poly(allylamine hydrochloride, Mw = 900,000, Sigma-Aldrich) was utilized as a surface modifier to form an amine group (NH2-) on the surface, inducing a positively charged surface, i.e., -NH3+-grafted MoS2@CNFs, in an aqueous solution. The MoS2@rGO was added to the PAH solution. After stirring for 2 h, the PAH-modified MoS2@rGO was rinsed three times with distilled water and dried at 60 °C in a vacuum oven overnight. Lastly, the modified sample was encapsulated with rGO according to the method in the same manner of our previous works [7,24], resulting in MoS2@CNFs@rGO.

2.2. Materials Characterization

Nova 230 (field-emission scanning electron microscope (FE-SEM), FEI, Hillsboro, OR, USA) was employed to obtain FE-SEM images. The crystal structure of MoS2@CNFs@rGO was investigated by X-ray diffraction (XRD) patterns using D/Max-2500, with RIGAKU Corp. (Tokyo, Japan) with Cu Kα (λ = 1.54 Å) between 10° and 80° at a scan rate of 0.066° s−1. Both internal and external morphologies of MoS2@CNFs@rGO and the distribution of elementals were analyzed by a high-resolution transmission electron microscope (HR-TEM) operating at 300 kV and a scanning TEM (STEM) using a Tecnai F30 S-Twin (FEI, Hillsboro, OR, USA) equipped with energy-dispersive X-ray spectroscopy (EDX). The chemical states of MoS2@CNFs@rGO were investigated by X-ray photoelectron spectroscopy (XPS, K-alpha, Thermo VG Scientific, Waltham, MA, USA). In addition, the dominant vibration modes in the MoS2@CNFs@rGO were investigated using Raman spectroscopy (ARAMIS, Horiba Jobin Yvon, Montpellier, France) with a 514 nm laser source.

2.3. Electrochemical Evaluation

All the electrodes were prepared by casting a slurry, including active materials (80%), a conducting agent (Super-P, Sigma-Aldrich, Burlington, MA, USA) (10%), and a polyvinylidene fluoride (PVDF, Mw ~534,000, Sigma-Aldrich, Burlington, MA, USA) binder (10%); the slurry was mixed together using an agate mortar and cast on a copper foil as a current collector. After casting, the electrode was dried in a vacuum oven for 12 h. The mass loading of the MoS2@CNFs@rGO was approximately 1.0 ± 0.1 mg cm−2. Half-cell (2032 type-coin cell) assembly was done in an Ar-filled glovebox (water content < 0.1 ppm). Na metal was used as a counter electrode, and a Whatman glass microfilter was employed as a separator for SIB cell tests. The used SIB electrolyte was 1 M NaClO4 in propylene carbonate (PC) with 5 wt % of FEC. The assembled coin cells for Na storge were cycled at a current density of 100~20,000 mA g1 between 0.005 and 3.0 V using a battery tester (WBCS3000 device by WonATech, Seoul, South Korea).

3. Results

3.1. Materials Characterization

The synthetic procedures to prepare the MoS2 nanoflakes confined in CNFs wrapped with the rGO net (MoS2@CNFs@rGO) are schematically illustrated in Figure 1. The MoS2-nanoflake-embedded carbon nanofibers (MoS2@CNFs) were prepared by thermal treatment using ammonium tetrathiomolybdate ((NH4)2MoS4) and poly(styrene-acrylonitrile) (SAN) as precursors through the electrospinning method (Figure 1a). First, the (NH4)2MoS4 was decomposed into MoS2 at 450 °C in the presence of reducing (H2) gas [25,26]. Then, carbonization and crystallization steps for MoS2 are performed through heat treatment at 800 °C under an inert gas environment. The reactions that occur during the heat treatment process are as follows.
( NH 4 ) 2 MoS 4   MoS 3 + 2 NH 3 +   H 2 S
MoS 3     MoS 2 +   S
( NH 4 ) 2 MoS 4 + H 2 MoS 2 + 2 NH 3 + 2 H 2 S
For rGO coating on the MoS2@CNFs, the surface of MoS2@CNFs was modified using poly(allylamine hydrochloride) (PAH) [7,23,24]. Firstly, the MoS2@CNFs were dipped in PAH solution; the surface functional group changed from the hydroxy surface group to the amine group (-NH2) (see the details in the Materials and Methods section). The PAH-modified MoS2@CNFs have positively charged amine functional groups (NH3+) exposed on the surface. On the other hand, the GO has functional groups of the carboxyl group (-COOH) and hydroxy group (-OH), attributed to the negative charge of GO in the aqueous solution. The GO flakes were quickly attracted to the surface of PAH-modified MoS2@CNFs due to the electrostatic self-assembly (Figure 1b). Then, the GO was chemically reduced by hydrazine treatment, which rendered the strong chemical bond between the amine functional group and oxygen group with the ring-opening reaction. Finally, the GO changed to rGO, covering the entire surface of MoS2@CNFs (Figure 1c).
The morphology of synthesized MoS2@CNFs is shown in Figure 2a,b, demonstrating that the nanocomposite possessed one-dimensional (1D) NF networks, with the diameter of each fiber at about 200 nm and several microsized pores among the NFs. These pores facilitate electrolyte penetration to enhance electrochemical performance. The morphological change of MoS2@CNFs@rGO was confirmed through SEM images (Figure 2c,d). It is confirmed that rGO flakes were covered on the NF surfaces while the 1D architecture of MoS2@CNFs was maintained. To verify the internal structure and phase of the MoS2@CNFs and MoS2@CNFs@rGO, transmission electron microscopy (TEM) analysis was conducted. For the MoS2@CNFs, few-layer MoS2 flakes were uniformly distributed into CNFs (Figure 3a).
After the graphene wrapping process, the rGO flakes substantially covered the MoS2@CNFs without any aggregations. Individual NFs were wrapped by stacked rGO flakes with a thickness of approximately 3 nm (Figure 3b,c). We expected that the favorable influx of Na+ into MoS2@CNFs would be possible through the thin rGO layer, and the electrically conductive rGO would enhance the electrical efficiency of the battery cell by facilitating the electron transport. Figure 3d shows a high-resolution TEM (HR-TEM) image of MoS2@CNFs@rGO. The lattice fringe of the MoS2 in the MoS2@CNFs@rGO is approximately 0.660 nm, indicating that the multilayer MoS2 nanoflakes had enlarged interlayer spacing compared to the general interlayer spacing (~0.615 nm) of multilayer MoS2 [27,28]. The enlarged interlayers can be attributed to the CNFs suppressing the crystallization of MoS2 during thermolysis. In addition, the wide interlayer distance might allow Na+ to diffuse rapidly, promoting Na+ flux in the whole electrode. The homogeneous distribution of elements—such as C (yellow), Mo (red), and S (green)—in the MoS2@CNFs@rGO was investigated by energy-dispersive spectroscopy (EDS) mapping images from a scanning TEM (Figure 3e). The elements were well-distributed in the 1D NF composite, which can be supported by the qualitative data in Figure S1.
To further consider the crystal structure and phase information of the MoS2@CNFs@rGO, Figure 4 exhibits the X-ray diffraction (XRD) patterns of MoS2@CNFs and MoS2@CNFs@rGO. At 14.2°, 33.3°, and 59.1°, the XRD peaks correspond to planes (002), (100), and (110) of the MoS2 phase (JCPDS #37-1492), respectively [29]. Through the graphene wrapping process, no peak shifts were observed for MoS2@CNFs@rGO. Furthermore, as a result of XPS analysis, the chemical states of Mo and C of MoS2@CNFs@rGO showed insignificant change despite the chemical reduction reaction. (Figure S2). Figure 5 compares the Raman spectra of MoS2@CNFs and MoS2@CNFs and shows the vibrational modes, the fingerprint of the chemical state of the MoS2 phase. MoS2@CNFs and MoS2@CNFs@rGO represent two peaks at 379.1 and 402.7 cm−1 due to in-plane E12g and out-of-plane A1g vibration modes [30]. Several studies exhibited that the relative intensity of these two peaks suggests the characteristics of MoS2 crystals, given dimensions and edge profiles. Generally, the intensity of the A1g mode is greater than that of the E12g mode when the MoS2 flakes have an edge-end structure. There is no difference between the E12g and A1g mode peaks for MoS2@CNFs and MoS2@CNFs@rGO, indicating that the MoS2 nanoflakes in the CNFs maintain their structures without damage during the fabrication process. Furthermore, introducing the rGO layer increases the intensity of the D and G bands, which indicates the carbon structure (Figure 5b and Figure S3) [31,32]. The integrated area ratio of sp3 to sp2 (Asp3/Asp2) has been proven to provide useful information concerning the nature of carbon [32]. The low ratio of Asp3/Asp2 indicates the presence of a large amount of sp2 carbon. The Asp3/Asp2 were 1.31 for MoS2@CNFs and 1.15 for MoS2@CNFs@rGO, respectively. It means that the amount of sp2-type carbon increased after rGO wrapping and the reduced graphene oxide layers are well-introduced in the MoS2@CNFs@rGO sample.
To quantitatively identify the contents of rGO in MoS2@CNFs@rGO, element analysis (EA) was carried out (Table 1). The MoS2@CNFs contain 23.6 wt % of carbon and about 29.8 wt % of sulfur from MoS2. After graphene wrapping, the amount of carbon (C) increased to 34.3 wt %, indicating that the sulfur (S) content was relatively decreased. Assuming that the ratio of CNF/MoS2 is maintained after graphene wrapping, it is confirmed that the content of rGO accounts for about 17.7 wt % in the MoS2@CNFs@rGO. Since the 1D nanostructure of MoS2@CNFs has a large surface area, a sufficient amount of rGO is required for the wrapping, even if it contains thin rGO layers. Therefore, the rGO should cover both the individual NFs and bundle of MoS2@CNFs, increasing the electronic conductivity of the MoS2@CNFs@rGO and preserving active materials, particularly S, during repetitive reactions [33,34].

3.2. Electrochemical Measerment

In general, MoS2 electrochemically reacts with Na+ based on insertion and conversion reactions [35]. To testify to these electrochemical behaviors in the MoS2@CNFs@rGO, the galvanostatic charge-discharge curves for the MoS2@CNFs@rGO were obtained in a voltage window between 0.005~3.0 V (vs. Na/Na+) at a current density of 100 mA g−1 (Figure 6a). The initial discharge and charge capacities of MoS2@CNFs@rGO are 1175 and 573 mAh g−1, respectively, corresponding to a Coulombic efficiency (CE) of 48%. This capacity fading for the initial cycle is attributed to the generation of high irreversible capacitance stemming from the solid-electrolyte interphase (SEI) layers formed on the electrode with a large surface area. Nevertheless, in Figure S4, the initial CE of MoS2@CNFs@rGO is higher than that of MoS2@CNFs (44%). It appears that more SEI layers seem to form on the MoS2@CNFs without the rGO coating layers while the rGO adequately stabilizes the SEI layers on the MoS2@CNFs@rGO [36]. Moreover, rGO played a critical role in accelerating electron transport, which improved the initial CE of MoS2@CNFs@rGO. It is confirmed that the MoS2@CNFs@rGO has a voltage plateau at 1.4 V (vs. Na/Na+) and a slope thereafter. This can be explained by a reaction caused by the insertion of Na+ into MoS2 (Equation (4)) and a subsequent conversion reaction (Equation (5)). Sodium polysulfide intermediate (Na2Sx, where x = 2 to 5) generated by the conversion reaction are easily dissolved in liquid electrolyte and move to the Na anode (“polysulfide shuttling”), leading to capacity loss and adverse effects on battery operation. The overall reaction can be represented as [37,38,39]:
MoS 2 + x   Na + + x   e   Na x MoS 2   ( x < 2 )
Na x MoS 2 + ( 4 x )   Na +   2   Na 2 S   + Mo
The voltage plateau generated at about 1.6 V in the charging process occurred due to the reduction of Na2S to S. After the initial cycle, the SIB cell containing the MoS2@CNFs@rGO electrode shows a highly reversible Na+ storage ability. We evaluated the long-term stability of MoS2@CNFs@rGO, as shown in Figure 6b. The MoS2@CNFs@rGO delivered a high discharge capacity of 345.8 mAh g−1 at the 90th cycle with a CE of 99.8%. In addition, the rate capability of MoS2@CNFs@rGO was tested at various current densities between 0.1~20 A g−1 (Figure 6c). At lower rates, the MoS2@CNFs@rGO exhibited similar capacity retention to that of MoS2@CNFs. On the other hand, the rate capability of MoS2@CNFs@rGO was gradually improved at higher rates (>0.5 A g−1). Moreover, the MoS2@CNFs@rGO outperformed the MoS2@CNFs even at a super-fast rate of 20 A g−1. This outstanding ability to store Na+ was also able to be used for Li+ storage (Figure S5).
All things taken together, we elucidated the reaction mechanism of MoS2@CNFs@rGO, based on the following factors: (I) MoS2 nanoflakes with enlarged interlayer spacing and short lateral distance between each flake are beneficial for storing Na+ in the CNFs. (II) The CNFs serve as a continuous passage for electrons and storage matrix for the active materials (Na+ and S). (III) The rGO is synergistically advantageous in terms of fast kinetics. Eventually, it turns out that the MoS2@CNFs@rGO synthesized by our fabrication approach enables the SIB cells to show remarkable performance toward Na+ insertion and conversion reactions.

4. Conclusions

In summary, we report a straightforward approach to fabricate a composite consisting of MoS2 nanoflakes confined in CNFs wrapped with a rGO net for SIB anodes. We effectively generated randomly distributed MoS2 nanoflakes (a few layers) in the CNFs via thermolysis and wrapped the rGO onto the MoS2@CNFs to enable the composite to be active for Na+ storage. The MoS2 nanoflakes possessing short lateral and expanded interlayer distances would be favorable for reversible insertion/desertion of Na+. In addition, the strongly interconnected CNFs and rGO net promoted electron transfer through the whole electrode, rendering a high rate of capability and cyclability. Furthermore, the rGO would preclude the dissolution of S stemming from the conversion reactions at the outermost surfaces of the composite materials. The MoS2@CNFs@rGO showed excellent cycle retention with a specific capacity of 345.8 mAh g−1 at a current density of 100 mA g−1 from the initial cycle to the 90th cycle. We elucidated that in the MoS2-based anodes undergoing conversion reactions, caging the active materials in the electronically conductive scaffold is critical for improving electrochemical battery performance.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/nano11102691/s1. Figure S1: EDS spectrum of rGO@MoS2@CNFs, Figure S2: XPS analysis for MoS2@CNFs@rGO and MoS2@CNFs and for (a) C1s and (b) Mo 3d, Figure S3: Raman analysis for (a) MoS2@CNFs@rGO and (b) MoS2@CNFs with classified peaks (I, IIII: sp3 bonding, II, IV: sp2 bonding), Figure S4: Charge–discharge curves of MoS2@CNFs at a current density of 100 mA g−1, Figure S5: Lithium-ion battery performance of MoS2@CNFs@rGO.

Author Contributions

Conceptualization, S.-H.C. and J.-W.J.; methodology, S.-H.C., J.-H.K. and I.-G.K.; validation, S.-H.C., I.-G.K. and J.-H.P.; formal analysis and investigation, S.-H.C. and J.-W.J.; writing—original draft preparation, S.-H.C., J.-H.K. and J.-W.J.; writing—review and editing as well as supervision, J.-W.J., H.-S.K. and I.-D.K.; project administration and funding acquisition, J.-W.J. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the 2021 Research Fund of the University of Ulsan.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Acknowledgments

This work was supported by the 2021 Research Fund of the University of Ulsan.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Senba, T. Clean Disruption of Energy and Transportation; Clean Planet Ventures: Silicon Valley, CA, USA, 2014; pp. 1–290. [Google Scholar]
  2. Cano, Z.P.; Banham, D.; Ye, S.; Hintennach, A.; Lu, J.; Fowler, M.; Chen, Z. Batteries and fuel cells for emerging electric vehicle markets. Nat. Energy 2018, 3, 279–289. [Google Scholar] [CrossRef]
  3. Tian, Y.; Zeng, G.; Rutt, A.; Shi, T.; Kim, H.; Wang, J.; Koettgen, J.; Sun, Y.; Ouyang, B.; Chen, T.; et al. Promises and Challenges of Next-Generation “Beyond Li-ion” Batteries for Electric Vehicles and Grid Decarbonization. Chem. Rev. 2021, 121, 1623–1669. [Google Scholar] [CrossRef]
  4. Usiskin, R.; Lu, Y.; Popovic, J.; Law, M.; Balaya, P.; Hu, Y.-S.; Maier, J. Fundamentals, status and promise of sodium-based batteries. Nat. Rev. 2021. [Google Scholar] [CrossRef]
  5. Hwan, J.-Y.; Myung, S.-T.; Sun, Y.-K. Sodium-ion batteries: Present and future. Chem. Soc. Rev. 2017, 46, 3529–3614. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Xiao, L.; Lu, H.; Fang, Y.; Sushko, M.L.; Cao, Y.; Ai, X.; Yang, H.; Liu, J. Low-Defect and Low-Porosity Hard Carbon with High Coulombic Efficiency and High Capacity for Practical Sodium Ion Battery Anode. Adv. Energy Mater. 2018, 8, 1703238. [Google Scholar] [CrossRef]
  7. Yeo, Y.; Jung, J.-W.; Park, K.; Kim, I.-D. Graphene-Wrapped Anatase TiO2 Nanofibers as High-Rate and Long-Cycle-Life Anode Material for Sodium Ion Batteries. Sci. Rep. 2015, 5, 13862. [Google Scholar] [CrossRef] [PubMed]
  8. Wu, L.; Shao, H.; Yang, C.; Feng, X.; Han, L.; Zhou, Y.; Du, W.; Sun, X.; Xu, Z.; Zhang, X.; et al. SnS2 Nanosheets with RGO Modification as High-Performance Anode Materials for Na-Ion and K-Ion Batteries. Nanomaterials 2021, 11, 1932. [Google Scholar] [CrossRef] [PubMed]
  9. Hu, L.; Shang, C.; Akinoglu, E.M.; Wang, X.; Zhou, G. Cu2Se Nanoparticles Encapsulated by Nitrogen-Doped Carbon Nanofibers for Efficient Sodium Storage. Nanomaterials 2020, 10, 302. [Google Scholar] [CrossRef] [PubMed]
  10. Kim, D.-Y.; Kim, D.-H.; Kim, S.-H.; Lee, E.-K.; Park, S.-K.; Lee, J.-W.; Yun, Y.-S.; Choi, S.-Y.; Kang, J. Hard Carbon Anodes for Sodium-Ion Batteries. Nanomaterials 2021, 9, 793. [Google Scholar] [CrossRef] [Green Version]
  11. Li, X.; Qi, S.-H.; Zhang, W.-C.; Feng, Y.-Z.; Ma, J.-M. Recent progress on FeS2 as anodes for metal-ion batteries. Rare Met. 2020, 39, 1239–1255. [Google Scholar] [CrossRef]
  12. Hu, R.; Fang, Y.; Liu, X.; Zhu, K.; Cao, D.; Yi, J.; Wang, G. Synthesis of SnS2 Ultrathin Nanosheets as Anode Materials for Potassium Ion Batteries. Chem. Res. Chin. Univ. 2021, 37, 311–317. [Google Scholar] [CrossRef]
  13. Yun, Q.; Li, L.; Hu, Z.; Lu, Q.; Chen, B.; Zhang, H. Layered Transition Metal Dichalcogenide-Based Nanomaterials for Electrochemical Energy Storage. Adv. Mater. 2019, 32, 1903826. [Google Scholar] [CrossRef]
  14. Yun, Q.; Lu, Q.; Zhang, X.; Tan, C.; Zhang, H. Three-Dimensional Architectures Constructed from Transition-Metal Dichalcogenide Nanomaterials for Electrochemical Energy Storage and Conversion. Angew. Chem. Int. Ed. 2017, 57, 626–646. [Google Scholar] [CrossRef]
  15. Wang, L.; Zhang, H.; Wang, Y.; Qian, C.; Dong, Q.; Deng, C.; Jiang, D.; Shu, M.; Pan, S.; Zhang, S. Unleashing ultra-fast sodium ion storage mechanisms in interface-engineered monolayer MoS2/C interoverlapped superstructure with robust charge transfer networks. J. Mater. Chem. A 2020, 8, 15002–15011. [Google Scholar] [CrossRef]
  16. Geng, X.; Jiao, Y.; Han, Y.; Mukhopadhyay, A.; Yang, L.; Zhu, H. Freestanding Metallic 1T MoS2 with Dual Ion Diffusion Paths as High Rate Anode for Sodium-Ion Batteries. Adv. Funct. Mater. 2017, 27, 1702998. [Google Scholar] [CrossRef]
  17. Jung, J.-W.; Ryu, W.-H.; Yu, S.; Kim, C.; Cho, S.-H.; Kim, I.-D. Dimensional Effects of MoS2 Nanoplates Embedded in Carbon Nanofibers for Bifunctional Li and Na Insertion and Conversion Reactions. ACS Appl. Mater. Interfaces 2016, 8, 26758–26768. [Google Scholar] [CrossRef]
  18. Li, Y.; Liang, Y.; Robles-Hernández, F.C.; Yoo, H.D. Enhancing sodium-ion battery performance with interlayer-expanded MoS2–PEO nanocomposites. Nano Energy 2015, 15, 453–461. [Google Scholar] [CrossRef] [Green Version]
  19. Ryu, W.-H.; Jung, J.-W.; Park, K.; Kim, S.-J.; Kim, I.-D. Vine-like MoS2 anode materials self-assembled from 1-D nanofibers for high capacity sodium rechargeable batteries. Nanoscale 2014, 6, 10975–10981. [Google Scholar] [CrossRef]
  20. Wang, L.; Yang, G.; Wang, J.; Peng, S.; Yan, W.; Ramakrishna, S. Controllable Design of MoS2 Nanosheets Grown on Nitrogen-Doped Branched TiO2/C Nanofibers: Toward Enhanced Sodium Storage Performance Induced by Pseudocapacitance Behavior. Small 2019, 16, 1904589. [Google Scholar] [CrossRef]
  21. Huang, Y.; Wang, Z.; Guan, M.; Wu, F.; Chen, R. Toward Rapid-Charging Sodium-Ion Batteries using Hybrid-Phase Molybdenum Sulfide Selenide-Based Anodes. Adv. Mater. 2020, 32, 2003534. [Google Scholar] [CrossRef]
  22. Kim, M.; Seok, H.; Selvam, N.C.S.; Cho, J.; Choi, G.H.; Nam, M.G.; Kang, S.; Kim, T.; Yoo, P.J. Kirkendall effect induced bifunctional hybrid electrocatalyst (Co9S8@MoS2/N-doped hollow carbon) for high performance overall water splitting. J. Power Source 2021, 493, 229688. [Google Scholar] [CrossRef]
  23. Zhou, W.; Zhu, J.; Cheng, C.; Liu, J.; Yang, H.; Cong, C.; Guan, C.; Jia, X.; Fan, H.J.; Yan, Q.; et al. A general strategy toward graphene@metal oxide core–shell nanostructures for high-performance lithium storage. Energy Environ. Sci. 2011, 4, 4954–4961. [Google Scholar] [CrossRef]
  24. Cho, S.-H.; Jung, J.-W.; Kim, C.; Kim, I.-D. Rational Design of 1-D Co3O4 Nanofibers@Low content Graphene Composite Anode for High Performance Li-Ion Batteries. Sci. Rep. 2017, 7, 45105. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Brio, J.L.; Ilija, M.; Hernández, P. Thermal and reductive decomposition of ammonium thiomolybdates. Thermochim. Acta 1995, 256, 325–338. [Google Scholar] [CrossRef]
  26. Müller, A.; Prasad, T.P.; Menge, R. Thermal decomposition of (NH4)2MoS4 and (NH4)2WS4 heat of formation of (NH4)2MoS4. Z. Für Anorg. Und Allg. Chem. 1972, 391, 107–112. [Google Scholar] [CrossRef]
  27. Wu, C.; Song, H.; Tang, C.; Du, A.; Yu, C.; Huang, Z.; Wu, M.; Zhang, H. Ultralarge interlayer distance and C,N-codoping enable superior sodium storage capabilities of MoS2 nanoonions. Chem. Eng. J. 2019, 378, 122249. [Google Scholar] [CrossRef]
  28. Sahu, T.S.; Mitra, S. Exfoliated MoS2 Sheets and Reduced Graphene Oxide-An Excellent and Fast Anode for Sodium-ion Battery. Sci. Rep. 2015, 5, 12571. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Yang, L.; Cui, X.; Zhang, J.; Wang, K.; Shen, M.; Zeng, S.; Dayeh, S.A.; Feng, L.; Xiang, B. Lattice strain effects on the optical properties of MoS2 nanosheets. Sci. Rep. 2014, 4, 5649. [Google Scholar] [CrossRef]
  30. Scheuschner, N.; Gillen, R.; Staiger, M.; Maultzsch, J. Interlayer resonant Raman modes in few-layer MoS2. Phys. Rev. B 2015, 91, 235409. [Google Scholar] [CrossRef] [Green Version]
  31. Wu, J.-B.; Lin, M.-L.; Cong, X.; Liu, H.-N.; Tan, P.-H. Raman spectroscopy of graphene-based materials and its applications in related devices. Chem. Soc. Rev. 2018, 47, 1822–1873. [Google Scholar] [CrossRef] [Green Version]
  32. Abdelaal, M.M.; Hung, T.-C.; Mohamed, S.G.; Yang, C.-C.; Huang, H.-P.; Hung, T.-F. A Comparative Study of the Influence of Nitrogen Content and Structural Characteristics of NiS/Nitrogen-Doped Carbon Nanocomposites on Capacitive Performances in Alkaline Medium. Nanomaterials 2021, 11, 1867. [Google Scholar] [CrossRef] [PubMed]
  33. Huang, Y.; Zhang, L.; Lu, H.; Lai, F.; Miao, Y.-E.; Liu, T. A highly flexible and conductive graphene-wrapped carbon nanofiber membrane for high-performance electrocatalytic application. Inorg. Chem. Front. 2016, 3, 969–976. [Google Scholar] [CrossRef]
  34. Cai, J.; Reinhart, B.; Eng, P.; Liu, Y.; Sun, C.-J.; Zhou, H.; Ren, Y.; Meng, X. Nitrogen-doped graphene-wrapped Cu2S as a superior anode in sodium-ion batteries. Carbon 2020, 170, 430–438. [Google Scholar] [CrossRef]
  35. Wang, R.; Wang, S.; Zhang, Y.; Jin, D.; Tao, X.; Zhang, L. Sodium storage in a promising MoS2–carbon anode: Elucidating structural and interfacial transitions in the intercalation process and conversion reactions. Nanoscale 2018, 10, 11165–11175. [Google Scholar] [CrossRef]
  36. Prosini, P.P.; Carewska, M.; Cento, C.; Tarquini, G.; Maroni, F.; Birrozzi, A.; Nobili, F. Tin-Decorated Reduced Graphene Oxide and NaLi0.2Ni0.25Mn0.75Oδ as Electrode Materials for Sodium-Ion Batteries. Materials 2019, 12, 1074. [Google Scholar] [CrossRef] [Green Version]
  37. Lei, T.; Chen, W.; Lv, W.; Huang, J.; Zhu, J.; Chu, J.; Yan, C.; Wu, C.; Yan, Y.; He, W.; et al. Inhibiting Polysulfide Shuttling with a Graphene Composite Separator for Highly Robust Lithium-Sulfur Batteries. Joule 2018, 2, 2091–2104. [Google Scholar] [CrossRef] [Green Version]
  38. Xu, X.; Zhou, D.; Qin, X.; Lin, K.; Kang, F.; Li, B.; Shanmukaraj, D.; Rojo, T.; Armand, M.; Wang, G. A room-temperature sodium–sulfur battery with high capacity and stable cycling performance. Nat. Commun. 2018, 9, 3870. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Mosa, J.; Garcia-Garcia, F.J.; González-Elipe, A.R.; Aparicio, M. New Insights on the Conversion Reaction Mechanism in Metal Oxide Electrodes for Sodium-Ion Batterie. Nanomaterials 2021, 11, 966. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic illustration for preparation of MoS2@CNFs@rGO. (a) Fabrication of MoS2@CNFs synthesized by thermolysis after electrospinning. (b) Electrostatic interaction between PAH-modified MoS2@CNFs and GO. (c) The final product of rGO-wrapped MoS2@CNFs achieved after hydrazine reduction.
Figure 1. Schematic illustration for preparation of MoS2@CNFs@rGO. (a) Fabrication of MoS2@CNFs synthesized by thermolysis after electrospinning. (b) Electrostatic interaction between PAH-modified MoS2@CNFs and GO. (c) The final product of rGO-wrapped MoS2@CNFs achieved after hydrazine reduction.
Nanomaterials 11 02691 g001
Figure 2. SEM images of (a) MoS2@CNFs and (b) MoS2@CNFs@rGO, HR-SEM images of (c) MoS2@CNFs and (d) MoS2@CNFs@rGO.
Figure 2. SEM images of (a) MoS2@CNFs and (b) MoS2@CNFs@rGO, HR-SEM images of (c) MoS2@CNFs and (d) MoS2@CNFs@rGO.
Nanomaterials 11 02691 g002
Figure 3. TEM images of (a) MoS2@CNFs and (b) MoS2@CNFs@rGO, (c,d) HR-TEM images of MoS2@CNFs@rGO, (e) STEM image and elemental distribution of C, Mo, and S in the MoS2@CNFs@rGO (scale bar: 20 nm).
Figure 3. TEM images of (a) MoS2@CNFs and (b) MoS2@CNFs@rGO, (c,d) HR-TEM images of MoS2@CNFs@rGO, (e) STEM image and elemental distribution of C, Mo, and S in the MoS2@CNFs@rGO (scale bar: 20 nm).
Nanomaterials 11 02691 g003
Figure 4. X-ray diffraction pattern of MoS2@CNFs and MoS2@CNFs@rGO.
Figure 4. X-ray diffraction pattern of MoS2@CNFs and MoS2@CNFs@rGO.
Nanomaterials 11 02691 g004
Figure 5. Results of Raman analyses of MoS2@CNFs and MoS2@CNFs@rGO with different ranges: (a) 200~500 cm−1, (b) 1000~2000 cm−1.
Figure 5. Results of Raman analyses of MoS2@CNFs and MoS2@CNFs@rGO with different ranges: (a) 200~500 cm−1, (b) 1000~2000 cm−1.
Nanomaterials 11 02691 g005
Figure 6. Electrochemical performances of MoS2@CNFs@rGO. (a) Charge–discharge profile at 100 mA g−1 between 0.005 V~3.0 V vs. Na/Na+. (b) Cycling performance at 100 mA g−1. (c) Rate capability with different current densities (0.1~20 A g−1).
Figure 6. Electrochemical performances of MoS2@CNFs@rGO. (a) Charge–discharge profile at 100 mA g−1 between 0.005 V~3.0 V vs. Na/Na+. (b) Cycling performance at 100 mA g−1. (c) Rate capability with different current densities (0.1~20 A g−1).
Nanomaterials 11 02691 g006
Table 1. Contents of the carbon and sulfur components of MoS2@CNFs and MoS2@CNFs@rGO.
Table 1. Contents of the carbon and sulfur components of MoS2@CNFs and MoS2@CNFs@rGO.
SamplesCarbon, C (wt %)Sulfur, S (wt %)
MoS2@CNFs23.629.8
MoS2@CNFs@ rGO34.320.9
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Cho, S.-H.; Kim, J.-H.; Kim, I.-G.; Park, J.-H.; Jung, J.-W.; Kim, H.-S.; Kim, I.-D. Reduced Graphene-Oxide-Encapsulated MoS2/Carbon Nanofiber Composite Electrode for High-Performance Na-Ion Batteries. Nanomaterials 2021, 11, 2691. https://doi.org/10.3390/nano11102691

AMA Style

Cho S-H, Kim J-H, Kim I-G, Park J-H, Jung J-W, Kim H-S, Kim I-D. Reduced Graphene-Oxide-Encapsulated MoS2/Carbon Nanofiber Composite Electrode for High-Performance Na-Ion Batteries. Nanomaterials. 2021; 11(10):2691. https://doi.org/10.3390/nano11102691

Chicago/Turabian Style

Cho, Su-Ho, Jong-Heon Kim, Il-Gyu Kim, Jeong-Ho Park, Ji-Won Jung, Hyun-Suk Kim, and Il-Doo Kim. 2021. "Reduced Graphene-Oxide-Encapsulated MoS2/Carbon Nanofiber Composite Electrode for High-Performance Na-Ion Batteries" Nanomaterials 11, no. 10: 2691. https://doi.org/10.3390/nano11102691

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