Next Article in Journal
The Effects of In Vivo Exposure to Copper Oxide Nanoparticles on the Gut Microbiome, Host Immunity, and Susceptibility to a Bacterial Infection in Earthworms
Next Article in Special Issue
Hole Transfer Layer Engineering for CdTe Nanocrystal Photovoltaics with Improved Efficiency
Previous Article in Journal
Plasma-Assisted Chemical Vapor Deposition of F-Doped MnO2 Nanostructures on Single Crystal Substrates
Previous Article in Special Issue
Noble Metal Nanoparticles Incorporated Siliceous TUD-1 Mesoporous Nano-Catalyst for Low-Temperature Oxidation of Carbon Monoxide
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

W2C/WS2 Alloy Nanoflowers as Anode Materials for Lithium-Ion Storage

Department of Chemical and Biological Engineering, Gachon University, Seongnam-si, Gyeonggi-do 13120, Korea
*
Author to whom correspondence should be addressed.
Nanomaterials 2020, 10(7), 1336; https://doi.org/10.3390/nano10071336
Submission received: 20 June 2020 / Revised: 3 July 2020 / Accepted: 6 July 2020 / Published: 9 July 2020

Abstract

:
Recently, composites of MXenes and two-dimensional transition metal dichalcogenides have emerged as promising materials for energy storage applications. In this study, W2C/WS2 alloy nanoflowers (NFs) were prepared by a facile hydrothermal method. The alloy NFs showed a particle size of 200 nm–1 μm, which could be controlled. The electrochemical performance of the as-prepared alloy NFs was investigated to evaluate their potential for application as lithium-ion battery (LIB) anodes. The incorporation of W2C in the WS2 NFs improved their electronic properties. Among them, the W2C/WS2_4h NF electrode showed the best electrochemical performance with an initial discharge capacity of 1040 mAh g−1 and excellent cyclability corresponding to a reversible capacity of 500 mAh g−1 after 100 cycles compared to that of the pure WS2 NF electrode. Therefore, the incorporation of W2C is a promising approach to improve the performance of LIB anode materials.

Graphical Abstract

1. Introduction

Two-dimensional (2D) materials such as graphene and transition metal chalcogenides show great potential for energy storage and conversion applications owing to their large surface area, high conductivity, and good physical and chemical stability [1,2,3,4,5,6,7,8,9,10,11,12]. Among these 2D materials, MXenes, which have been discovered recently, have been extensively investigated for energy storage applications. It should be noted that MXenes are the transition metal carbides/nitride, which have graphene-like structure, possessing many advantages of 2D materials such as high conductivity, flexibility, easy processing, and so on [13]. Mashtalir et al. synthesized intercalated Ti3C2 flakes with high stability and the charging rate for application as an anode material for lithium-ion batteries (LIBs) [14]. Naguib et al. prepared niobium and vanadium carbide LIB anodes with high rate capacity [15]. The MXene anodes exhibited excellent electrochemical performance at high currents because of their low diffusion barrier [16]. In addition, Fe3O4@Ti2C3 LIB anodes exhibited an ultrahigh capacity of approximately 747 mAh g−1 at 1 C for 1000 cycles. Moreover, these materials exhibited a capacity of approximately 278 mAh g−1 at the high rate of 5 C [17]. Zhang et al. have demonstrated the use of MXenes as conductive binders for viscous aqueous inks of silicon materials, which were used as high-capacity anode materials for LIBs [18]. In these materials, MXenes not only acted as a conductive network for Si particles, but also improved the mechanical stability of the material.
Tungsten metal compounds such as oxides, chalcogenides, and carbides are used in a wide range of applications such as catalysis, energy conversion, and energy storage [19,20,21,22,23,24,25,26,27]. Feng et al. prepared WS2 nanoflakes with a high reversible capacity of 680 mAh g−1 for 20 cycles as anode materials for LIBs [27]. Srinivaas et al. have prepared highly rich 1T WS2 phase in few layered nanoflowers (NFs) with stable electrochemical performance and a high initial capacity of approximately 890 mAh g−1 as anode materials for LIBs [28]. W2C is a good catalyst. When grown on carbon nanotubes, it exhibits high photo/electrocatalytic performance for the hydrogen evolution reaction [26,29]. Simulation results have shown that W2C can exhibit an ultra-fast loading of lithium ions owing to its low diffusion barrier of about 0.045–0.13 eV [30]. Thus, the combination of WS2 and W2C is expected to yield a promising LIB anode material.
In this study, we prepared a W2C/WS2 NF composite via a facile hydrothermal process. By controlling the reaction time, alloy NFs with different sizes could be prepared. The structures, chemical compositions, and binding states of the as-prepared NFs were investigated. Finally, the potential of the W2C/WS2 NFs for LIB anode applications was investigated. The NFs showed promising electrochemical performance and excellent Li storage.

2. Materials and Methods

2.1. Chemical Materials

Tungsten (VI) chloride (WCl6, 99.9% trace metals basis, Sigma-Aldrich Inc., St. Louis, MO, USA), thioacetamide (C2H5NS, 99%, Sigma-Aldrich, Sigma-Aldrich Inc., St. Louis, MO, USA), amorphous carbon black Super-P (C, approximately 40 nm 99.99%, Alfa Aesar Inc., MA, USA), absolute ethanol (C2H5OH, Alpha Aesar Inc., MA, USA), and polyvinylidene fluoride (PVDF, 534,000 MW, Sigma-Aldrich, Sigma-Aldrich Inc., St. Louis, MO, USA) were used as received without any treatment.

2.2. Synthesis of W2C/WS2 NFs

The W2C/WS2 alloy NFs were prepared according to a previously reported procedure [31]. Briefly, 0.6 g of WCl6 was dispersed as the W source in a 20 mL tube with 4 mL of absolute ethanol. For the sulfur and carbon source, 1.2 g of thioacetamide was dispersed in 4 mL of absolute ethanol under stirring. Then, the WCl6 solution was quickly added to the thioacetamide solution, and the resulting solution was stirred for 5 min. Then, 30 mL of deionized (DI) water was added, and the solution was stirred for 1 h. Then, the reaction mixture was transferred to a 50 mL polypropylene-lined autoclave and heated in an oven at 250 °C for 2, 4, and 12 h to obtain the W2C/WS2_2h, W2C/WS2_4h, and W2C/WS2_12h precipitates, respectively. These precipitates were washed four times with ethanol and DI water and were then dried in an oven at 60 °C to obtain the W2C/WS2 powders.

2.3. Characterization

X-ray diffraction (XRD) (D/MAX-2200 Rigaku, Tokyo, Japan) was used to analyze the structure of the samples. The XRD patterns of the samples were recorded over the 2θ range of 5–70°. The morphologies, structures, and sizes of the samples were investigated by scanning electron microscopy (SEM) (Hitachi S4700, Tokyo, Japan) and transmission electron microscopy (TEM) (TECNAI G2F30, FEI corp., OR, USA). The Raman spectra of the samples were acquired using a Raman spectrometer (Lab RAM HR, Horiba JobinYvon, Horiba Ltd., Kyoto, Japan, 532 nm laser excitation). X-ray photoelectron spectroscopy (XPS) (Axis Ultra DLD, Kratos Analytical Ltd, Kyoto, Japan) under a high vacuum of 1.6 × 10−10 mbar with a monochromatic Al Kα line was used to investigate the chemical compositions and atomic binding of the samples.

2.4. Electrochemical Measurements

The LIBs were assembled using coin-type cells (CR 2032, Rotech Inc., Gwangju, Korea). The working electrode was prepared by casting a slurry of 70% active material (W2C/WS2 alloy NFs), 15% conductive carbon black (Super-P), and 15% PVDF in N-methyl-2-pyrrolidinone on a copper foil by doctor blading. After drying in a vacuum oven at 70 °C for 12 h, the electrodes were punched into circular discs with a diameter of 12 mm. The battery half-cell structures were assembled under an Ar2 atmosphere in a glovebox. A lithium foil, polyethylene, and 1M LiPF6 in ethylene carbonate/diethylene carbonate (1:1 in volume) were employed as the reference electrode, separator, and electrolyte, respectively. Galvanostatic electrochemical charge–discharge measurements were carried out using a battery cycle tester (WBCS3000, WonAtech, Seocho-gu, Seoul) over the voltage range of 0.1–3.0 V versus Li/Li+. Cyclic voltammetry (CV) tests were carried out using ZIVE MP1 (WonAtech, Seocho-gu, Seoul) over the voltage range of 0.1–3.0 V at a scanning rate of 0.1 mV s−1. Electrochemical impedance spectroscopic (EIS) measurements were carried out using a ZIVE MP1 (WonAtech, Seocho-gu, Seoul) over the frequency range of 100 kHz–0.1 Hz.

3. Results

For the synthesis of the W2C/WS2 NFs, the WCl6 precursor was pre-mixed with absolute ethanol (WCl6 + xC2H5OH → WCl6-x(OC2H5)x + xHCl) before mixing with thioacetamide to prevent its unexpected reaction with moisture (WCl6 + xH2O → WCl6-x(OH)x + xHCl) [31,32]. The hydrolysis of thioacetamide produced hydrosulfide, acetic acid, and ammonia (2C3H5NS + 6H2O → 2H2S + 3CH3COOH + 2NH3). The introduction of hydrosulfide, acetic acid, and ethanol, respectively, acted as the sulfur and carbon sources for the formation of the W2C/WS2 alloys. Figure 1a–c show the SEM images of the W2C/WS2_2h, W2C/WS2_4h, and W2C/WS2_12h NFs, respectively. The particle sizes of the W2C/WS2_2h, W2C/WS2_4h, and W2C/WS2_12h NFs were 200, 400, and 1000 nm, respectively. The NFs showed a highly uniform structure, indicating that the synthesis procedure was highly reproducible. Each NF consisted of several leaves of W2C nanosheets and WS2 nanocrystals. Meanwhile, the WS2 NFs synthesized without ethanol showed a complex structure because of the non-uniform dispersion of W (Figure 1d).
To further examine the structure of the W2C/WS2 NFs, powder XRD measurements were carried out over the 2θ range of 5–70° (Figure 2a). The XRD peaks of all the samples could be indexed to the hexagonal structure of W2C and WS2 [32,33,34,35]. It should be noted that the W2C/WS2_2h sample showed weak WS2 peaks. Moreover, this sample showed broad W2C peaks, indicating the small crystallite size of W2C. The W2C/WS2_4h and W2C/WS2_12h samples showed clear W2C and WS2 peaks, indicating the co-existence of these hexagonal-structured materials. For comparison, a WS2 sample was prepared using the same procedure as that used for preparing the W2C/WS2 NFs but without the addition of ethanol. This sample showed peaks corresponding to WS2 only. Therefore, the addition of ethanol during the synthesis not only prevented the oxidation of WCl6 but also contributed to the formation of carbide from the thioacetamide source. The structures of W2C and WS2 in the alloy NFs were examined by Raman spectroscopy (Figure 2b). The samples for Raman spectroscopy measurements were prepared on SiO2/Si substrates, and a 532-nm wavelength laser source was used. The samples showed two peaks at 351.5 and 415.2 cm−1 corresponding to the in-plane mode vibration E12g and out-plane mode vibration A1g of WS2, respectively [2,36]. Moreover, the peaks detected at 700 and 805 cm−1 can be attributed to the stretching modes of W−C [26,37]. The D- and G-bands corresponding to the sp3 and sp2 carbon atoms in W2C were observed at 1200–1700 cm−1. These results indicate the co-existence of W2C and WS2 in the alloy NFs.
The structure of the W2C/WS2 alloy NFs was further examined using TEM. As can be observed from Figure 3a, the samples showed a uniform NF structure with many leaves. Some Moiré patterns with a size of about 0.9–1.0 nm could be observed in the TEM images of the samples (Figure 3b). This can be attributed to the formation of the W2C/WS2 interlayer spacing [31]. The lattice spacings of W2C and WS2 in the samples were also measured from their high-resolution TEM (HRTEM) image shown in Figure 3c. The distance spacings of the WS2 (100) and W2C (101) planes were measured to be 0.27 and 0.23 nm, respectively. These spacings are characteristic of these materials. Moreover, the selected-area electron diffraction (SAED) pattern shown in Figure 3d indicates that W2C/WS2 alloy NFs with a highly crystalline hexagonal structure were successfully fabricated.
The chemical compositions and atomic binding energies of the W2C/WS2 alloys were analyzed by XPS (Figure 4a–d). The survey scan XPS profiles of the alloys showed clear and sharp peaks corresponding to W, S, O, C, and Si (from substrate) (Figure 4a). No impurity peak was detected, indicating the high purity of the W2C/WS2 alloys. The high-resolution W 4f, S 2p, and C 1s spectra of the alloys are shown in Figure 4b–d, respectively. The W 4f peak consisted of doublet peaks corresponding to the W−C, W−S, and W−O bonds. The contribution of W−O bonding can be attributed to the slight oxidation on the surface of the alloys, which always occurs during the preparation or natural oxidation of a material in air [29]. The W 4f7/2 and W 4f5/2 peaks of the W−C doublet were observed at 31.9 and 34.1 eV, respectively, while those of the W−S doublet were observed at 32.4 and 34.6 eV, respectively. The W 4f7/2 and W 4f5/2 peaks of the W−O doublet were observed at 36.0 and 38.2 eV, respectively. The S 2p peak of the samples could be deconvoluted into the S 2p3/2 and S 2p1/2 peaks, which were observed at 161.4 and 162.6 eV, respectively. Furthermore, the slight oxidation of the surface of the alloys during the preparation resulted in the appearance of the S−O bond peak at about 169 eV. Finally, the C 1s peak of the samples could be deconvoluted into those corresponding to C−W bonding at 284.7 eV, C−O bonding at 286.2 eV, and C−OH bonding at 288.5 eV. This is consistent with the binding energy of C in carbide compounds [38].
To evaluate the electrochemical performance of the alloys for application as LIB anode materials, their CV tests were carried out for three cycles over the voltage range of 0.1–3.0 V (Figure 5a–d). In the first cycle, the as-prepared WS2NFs showed lithiation at 0.5, 1.2, and 1.4 V attributing to the reduction of WS2 to Li2S through multiple steps, including the formation of LixWS2 and Li2S [28]. The peak observed at 0.5 V corresponds to the conversion reaction (4Li+ + WS2 + 4e → 2Li2S + W) as well as the decomposition of the non-aqueous electrolyte to form the solid electrolyte interface (SEI) layer [28,39]. In the second cycle, the peak at 0.5 V disappeared because of the formation of the SEI and gel-like polymeric layers by the dissolution of Li2S into the electrolyte, leading to its degradation [40]. The peaks at 1.2 and 1.4 V shifted to 1.27 V. An additional peak was observed at 1.9 V, indicating the insertion of Li into LixWS2 [28]. Figure 5b–d show the cyclic voltammograms of the W2C/WS2_2h, W2C/WS2_4h, and W2C/WS2_12h samples, respectively. In the first cycle, the W2C/WS2_2h sample showed a small peak at about 1.1 V. In the second and third cycles, the sample showed reduction peaks at 1.25 and 1.9 V attributing to the multi-step lithiation of WS2. Interestingly, the W2C/WS2_4h and W2C/WS2_12h samples with large NFs showed only a broad peak at 1.3–1.5 V in the first discharge cycle. This phenomenon can be attributed to the change in the work function of WS2 by the addition of W2C [31]. The changes in the work function of the alloys corresponded to the changes in their reduction potentials. The work function of the W2C/WS2 NFs increased from 4.31 to 4.7 eV with an increase in the reaction time from 2 to 12 h. The work function of the pure WS2 sample was 4.95 eV. The bare WS2 electrode showed lithiation peaks at 1.4 and 1.9 V. On the other hand, the W2C/WS2_2h and W2C/WS2_4h electrodes showed lithiation peaks at lower potentials at around 1.2–1.5 V. These samples did not show any lithiation peak at 1.9 V. Hence, it can be stated that the decrease in the work function resulted in a decrease in the lithiation potential of the samples. As shown by the XRD patterns, the W2C/WS2_2h sample showed broader W2C peaks than the other samples, indicating that the W2C crystallite size of this sample was smaller than those of the other samples. Therefore, the change in the second lithiation potential of the W2C/WS2_2h electrode was comparable to the initial potential of the WS2 electrode because of the instability of its smaller W2C crystals. The W2C/WS2_4h and W2C/WS2_12h electrodes only showed reduction peaks at about 1.2–1.5 V. This indicates that with an increase in the reaction time, the bonding between W2C and WS2 became stronger. The W2C/WS2 and WS2 NFs showed similar oxidation peaks because of the restoration of the WS2 structure at about 1.7–1.9 V and the oxidation of Li2S (Li2S → 2Li+S) at about 2.3–2.5 V [39]. The W2C/WS2 NFs showed an oxidation peak at approximately 1.2 V, which was attributed to the delithiation of W2C.
The initial discharge and charge voltage profiles of the W2C/WS2 and WS2 NFs were obtained over the voltage range of 0.1–3.0 V at a scan rate of 100 mA g−1, as shown in Figure 6a. The WS2 NFs showed the charge and discharge capacities of 504.0 and 656.6 mAh g−1, respectively. The W2C/WS2_2h, W2C/WS2_4h, and W2C/WS2_12h NFs showed the charge and discharge capacities of 595.2 and 935.2 mAh g−1, 751.8 and 1040.5 mAh g−1, 717.7 and 953.5 mAh g−1, respectively. It has been reported that carbide materials are promising candidates for energy storage applications [41]. In this study, the addition of W2C to the WS2 NF sample improved its storage capacity. The cyclic performances of the W2C/WS2 and WS2 NFs were evaluated over 100 cycles, as shown in Figure 6b. The alloys NF electrodes exhibited different electrochemical properties depending on the reaction time. For instance, after 10 cycles, the capacities of the W2C/WS2_2h and W2C/WS2_12h electrodes decayed rapidly by about 55%. After 30–40 cycles, the samples showed low stable capacities of 100–200 mAh g−1 with a high Coulombic efficiency of approximately 95–99%. On the other hand, the W2C/WS2_4h electrode showed excellent cyclic stability with a high capacity of 500 mAh g−1 after 100 cycles as compared to the other electrodes. The reasons why the W2C/WS2_4h electrode demonstrated the best performance are the stable binding of W2C to WS2 and the lowering of reduction potential in addition to the low resistance (discussed later). This sample showed a Coulombic efficiency of 97–98%. Meanwhile, the WS2 NFs showed a Coulombic efficiency of 93–94% and a continuous decrease in the cyclic capacity to approximately 145 mAh g−1 during the 100th cycle.
To further investigate the performance of the NFs, their EIS measurements were carried out over the frequency range of 100 kHz–0.1 Hz at a voltage of 3.0 V. Figure 7 shows the Nyquist plots of the WS2 and W2C/WS2 NF electrodes with the equivalent circuit as an inset containing the constant phase elements (CPE1, 2), series resistance (Rs), charge transfer resistance (Rct), solid electrolyte resistance (RSEI), and diffusion Warburg element (W). The semicircular arc corresponds to their charge transfer resistances between the interface of electrode and electrolyte [42]. The resistance values are summarized in Table 1. The WS2 NF anode showed the highest charge transfer resistance among all the anodes investigated. The charge transfer resistance of the WS2 NFs decreased with the addition of W2C in the case of the W2C/WS2_2h and W2C/WS2_4h electrodes. However, it increased again for the W2C/WS2_12h electrode. The difference in the performance of the as-prepared W2C/WS2 electrodes could be due to the different amount of W2C in the alloys. As shown by the Raman spectrum in Figure 2b, the peak intensities of the D- and G-band from the carbon atoms in W2C decreased when the reaction time increased, illustrating an increase in the amount of WS2. In the case of the W2C/WS2_12h sample, it could have an excess amount of H2S at 250 °C, and the applied high pressure sulfurized the weak bonding of W−C, leading to the lower amount of C, which was confirmed by Raman spectra representing the decrease of the D-band and G-band. Moreover, the flower size of the W2C/WS2_12h sample is about 1 μm, which is even bigger than the random size of WS2 NFs. It should be noted that the increase in size could lead to the decrease in the conductivity [28]. When the amount and size of WS2 increase, the resistance of the electrodes increases, as discussed in the EIS results. Specifically, the W2C/WS2_2h and 4h electrodes showed lower resistances compared to that of the W2C/WS2_12h electrode. Even though the W2C/WS2_2h electrode showed the lowest resistance value, the cycling performance was not stable due to the weak binding of the W2C crystal to WS2, leading to the gradual capacity decay, as discussed earlier. Thus, it can be concluded that the control of reaction time for the preparation of W2C/WS2 electrodes is crucial to optimize the overall electrochemical properties, where the W2C/WS24h alloy is the best electrode. Computational calculations have revealed that W2C materials are promising candidates for Li storage applications [30]. Li ions tend to adsorb on W2C materials to form a metal cluster. Moreover, MXenes nanosheets materials exhibit high conductivity [43]. In this study, the addition of W2C reduced the charge transfer resistance of the WS2 NFs, and the resulting W2C/WS2 alloy NFs showed high storage capacity and stability for LIB applications.

4. Conclusions

W2C/WS2 NFs were successfully fabricated via a facile hydrothermal method at low temperature. The particle size of the NFs could be controlled (200 nm–1 μm). The obtained NFs exhibited high purity and well-defined hexagonal structures of W2C and WS2. The NF alloys were employed as anode materials for LIBs. The W2C/WS2_4h sample showed a high initial discharge capacity of 1040 mAh g−1. The alloy electrodes showed a low charge transfer resistance of 200–600 Ω, indicating that the W2C/WS2 alloy electrodes were highly conductive as compared to the WS2 NF electrode. Among all the electrodes investigated, the W2C/WS2_4h electrode exhibited the highest stable capacity of up to approximately 500 mAh g−1 over 100 cycles, which could be attributed to the optimized W2C in WS2 NFs, Therefore, the W2C/WS2 alloy NFs prepared in this study showed a potential for energy storage and conversion applications.

Author Contributions

Conceptualization, T.P.N.; methodology, T.P.N.; validation, T.P.N.; writing—review and editing, T.P.N. and I.T.K.; visualization, T.P.N.; project administration, I.T.K.; funding acquisition, I.T.K. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Korea Electric Power Corporation (grant number: R18XA02). This research was supported by the Basic Science Research Capacity Enhancement Project of Korea Basic Science Institute (National research Facilities and Equipment Center) and was funded by the Ministry of Education. (2019R1A6C1010016).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Xu, M.; Liang, T.; Shi, M.; Chen, H. Graphene-Like Two-Dimensional Materials. Chem. Rev. 2013, 113, 3766–3798. [Google Scholar] [CrossRef] [PubMed]
  2. 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. Engl. 2010, 49, 4059–4062. [Google Scholar] [CrossRef]
  3. Kolobov, A.V.; Tominaga, J. Two-Dimensional Transition-Metal. Dichalcogenides; Springer International Publishing: Cham, Switzerland, 2016. [Google Scholar]
  4. Kalambate, P.K.; Gadhari, N.S.; Li, X.; Rao, Z.; Navale, S.T.; Shen, Y.; Patil, V.R.; Huang, Y. Recent advances in MXene–based electrochemical sensors and biosensors. TrAC Trends Anal. Chem. 2019, 120, 115643. [Google Scholar] [CrossRef]
  5. Naguib, M.; Mochalin, V.N.; Barsoum, M.W.; Gogotsi, Y. 25th Anniversary Article: MXenes: A New Family of Two-Dimensional Materials. Adv. Mater. 2014, 26, 992–1005. [Google Scholar] [CrossRef]
  6. Nguyen, T.P.; Tuan Nguyen, D.M.; Tran, D.L.; Le, H.K.; Vo, D.-V.N.; Lam, S.S.; Varma, R.S.; Shokouhimehr, M.; Nguyen, C.C.; Le, Q.V. MXenes: Applications in electrocatalytic, photocatalytic hydrogen evolution reaction and CO2 reduction. Mol. Catal. 2020, 486, 110850. [Google Scholar] [CrossRef]
  7. Gatensby, R.; McEvoy, N.; Lee, K.; Hallam, T.; Berner, N.C.; Rezvani, E.; Winters, S.; O’Brien, M.; Duesberg, G.S. Controlled synthesis of transition metal dichalcogenide thin films for electronic applications. Appl. Surf. Sci. 2014, 297, 139–146. [Google Scholar] [CrossRef]
  8. Phung, V.-D.; Jung, W.-S.; Kim, J.-H.; Lee, S.-W. Gold nanostructures electrodeposited on graphene oxide-modified indium tin oxide glass as a surface-enhanced Raman scattering-active substrate for ultrasensitive detection of dopamine neurotransmitter. Jpn. J. Appl. Phys. 2018, 57, 08PF02. [Google Scholar] [CrossRef]
  9. Nguyen, Q.H.; Hur, J. MoS2-C-TiC Nanocomposites as New Anode Materials for High-Performance Lithium-Ion Batteries. J. Nanosci. Nanotechnol. 2019, 19, 996–1000. [Google Scholar] [CrossRef]
  10. Nguyen, Q.H.; Nguyen, Q.H.; Hur, J. High-Performance ZnTe-TiO2-C nanocomposite with half-cell and full-cell applications as promising anode material for Li-Ion batteries. Appl. Surf. Sci. 2020, 509, 144718. [Google Scholar] [CrossRef]
  11. Liu, N.; Choi, W.; Kim, H.; Jung, C.; Kim, J.; Choo, S.H.; Kwon, Y.; An, B.-S.; Hong, S.; So, S.; et al. Rapid and mass-producible synthesis of high-crystallinity MoSe2 nanosheets by ampoule-loaded chemical vapor deposition. Nanoscale 2020, 12, 6991–6999. [Google Scholar] [CrossRef] [PubMed]
  12. Tran, A.-V.; Shim, K.; Vo Thi, T.-T.; Kook, J.-K.; An, S.S.A.; Lee, S.-W. Targeted and controlled drug delivery by multifunctional mesoporous silica nanoparticles with internal fluorescent conjugates and external polydopamine and graphene oxide layers. Acta Biomater. 2018, 74, 397–413. [Google Scholar] [CrossRef] [PubMed]
  13. Naguib, M.; Kurtoglu, M.; Presser, V.; Lu, J.; Niu, J.; Heon, M.; Hultman, L.; Gogotsi, Y.; Barsoum, M.W. Two-Dimensional Nanocrystals Produced by Exfoliation of Ti3AlC2. Adv. Mater. 2011, 23, 4248–4253. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Mashtalir, O.; Naguib, M.; Mochalin, V.N.; Dall’Agnese, Y.; Heon, M.; Barsoum, M.W.; Gogotsi, Y. Intercalation and delamination of layered carbides and carbonitrides. Nat. Commun. 2013, 4, 1–7. [Google Scholar] [CrossRef] [PubMed]
  15. Naguib, M.; Halim, J.; Lu, J.; Cook, K.M.; Hultman, L.; Gogotsi, Y.; Barsoum, M.W. New Two-Dimensional Niobium and Vanadium Carbides as Promising Materials for Li-Ion Batteries. J. Am. Chem. Soc. 2013, 135, 15966–15969. [Google Scholar] [CrossRef]
  16. Tang, Q.; Zhou, Z.; Shen, P. Are MXenes Promising Anode Materials for Li Ion Batteries? Computational Studies on Electronic Properties and Li Storage Capability of Ti3C2 and Ti3C2X2 (X = F, OH) Monolayer. J. Am. Chem. Soc. 2012, 134, 16909–16916. [Google Scholar] [CrossRef]
  17. Wang, Y.; Li, Y.; Qiu, Z.; Wu, X.; Zhou, P.; Zhou, T.; Zhao, J.; Miao, Z.; Zhou, J.; Zhuo, S. Fe3O4@Ti3C2 MXene hybrids with ultrahigh volumetric capacity as an anode material for lithium-ion batteries. J. Mater. Chem. A 2018, 6, 11189–11197. [Google Scholar] [CrossRef]
  18. Zhang, C.; Park, S.-H.; Seral-Ascaso, A.; Barwich, S.; McEvoy, N.; Boland, C.S.; Coleman, J.N.; Gogotsi, Y.; Nicolosi, V. High capacity silicon anodes enabled by MXene viscous aqueous ink. Nat. Commun. 2019, 10, 1–9. [Google Scholar] [CrossRef] [Green Version]
  19. Nayak, A.K.; Pradhan, D. Microwave-Assisted Greener Synthesis of Defect-Rich Tungsten Oxide Nanowires with Enhanced Photocatalytic and Photoelectrochemical Performance. J. Phys. Chem. C 2018, 122, 3183–3193. [Google Scholar] [CrossRef]
  20. Meyer, J.; Hamwi, S.; Kröger, M.; Kowalsky, W.; Riedl, T.; Kahn, A. Transition Metal Oxides for Organic Electronics: Energetics, Device Physics and Applications. Adv. Mater. 2012, 24, 5408–5427. [Google Scholar] [CrossRef]
  21. Emin, S.; Altinkaya, C.; Semerci, A.; Okuyucu, H.; Yildiz, A.; Stefanov, P. Tungsten carbide electrocatalysts prepared from metallic tungsten nanoparticles for efficient hydrogen evolution. Appl. Catal. B 2018, 236, 147–153. [Google Scholar] [CrossRef]
  22. Choi, H.; Kim, B.; Ko, M.J.; Lee, D.-K.; Kim, H.; Kim, S.H.; Kim, K. Solution processed WO3 layer for the replacement of PEDOT:PSS layer in organic photovoltaic cells. Org. Electron. 2012, 13, 959–968. [Google Scholar] [CrossRef]
  23. Li, W.-J.; Fu, Z.-W. Nanostructured WO3 thin film as a new anode material for lithium-ion batteries. Appl. Surf. Sci. 2010, 256, 2447–2452. [Google Scholar] [CrossRef]
  24. Chen, T.-Y.; Chang, Y.-H.; Hsu, C.-L.; Wei, K.-H.; Chiang, C.-Y.; Li, L.-J. Comparative study on MoS2 and WS2 for electrocatalytic water splitting. Int. J. Hydrogen Energy 2013, 38, 12302–12309. [Google Scholar] [CrossRef]
  25. Lukowski, M.A.; Daniel, A.S.; English, C.R.; Meng, F.; Forticaux, A.; Hamers, R.J.; Jin, S. Highly active hydrogen evolution catalysis from metallic WS2 nanosheets. Energy Environ. Sci. 2014, 7, 2608–2613. [Google Scholar] [CrossRef]
  26. Chen, Z.; Gong, W.; Cong, S.; Wang, Z.; Song, G.; Pan, T.; Tang, X.; Chen, J.; Lu, W.; Zhao, Z. Eutectoid-Structured WC/W2C heterostructures: A new platform for long-term alkaline hydrogen evolution reaction at low overpotentials. Nano Energy 2020, 68, 104335. [Google Scholar] [CrossRef]
  27. Feng, C.; Huang, L.; Guo, Z.; Liu, H. Synthesis of tungsten disulfide (WS2) nanoflakes for lithium ion battery application. Electrochem. Commun. 2007, 9, 119–122. [Google Scholar] [CrossRef]
  28. Srinivaas, M.; Wu, C.-Y.; Duh, J.-G.; Wu, J.M. Highly Rich 1T Metallic Phase of Few-Layered WS2 Nanoflowers for Enhanced Storage of Lithium-Ion Batteries. ACS Sustain. Chem. Eng. 2019, 7, 10363–10370. [Google Scholar] [CrossRef]
  29. Gong, Q.; Wang, Y.; Hu, Q.; Zhou, J.; Feng, R.; Duchesne, P.N.; Zhang, P.; Chen, F.; Han, N.; Li, Y.; et al. Ultrasmall and phase-pure W2C nanoparticles for efficient electrocatalytic and photoelectrochemical hydrogen evolution. Nat. Commun. 2016, 7, 1–8. [Google Scholar] [CrossRef] [Green Version]
  30. Zhang, Y. First principles prediction of two-dimensional tungsten carbide (W2C) monolayer and its Li storage capability. Comput. Condens. Matter 2017, 10, 35–38. [Google Scholar] [CrossRef]
  31. Nguyen, T.P.; Choi, K.S.; Kim, S.Y.; Lee, T.H.; Jang, H.W.; Van Le, Q.; Kim, I.T. Strategy for controlling the morphology and work function of W2C/WS2 nanoflowers. J. Alloys Compd. 2020, 829, 154582. [Google Scholar] [CrossRef]
  32. Nguyen, T.P.; Kim, S.Y.; Lee, T.H.; Jang, H.W.; Le, Q.V.; Kim, I.T. Facile synthesis of W2C@WS2 alloy nanoflowers and their hydrogen generation performance. Appl. Surf. Sci. 2020, 504, 144389. [Google Scholar] [CrossRef]
  33. Rout, C.S.; Joshi, P.D.; Kashid, R.V.; Joag, D.S.; More, M.A.; Simbeck, A.J.; Washington, M.; Nayak, S.K.; Late, D.J. Superior field emission properties of layered WS2-RGO nanocomposites. Sci. Rep. 2013, 3, 1–8. [Google Scholar] [CrossRef] [PubMed]
  34. Kurlov, A.S.; Gusev, A.I. Tungsten carbides and W-C phase diagram. Inorg. Mater. 2006, 42, 121–127. [Google Scholar] [CrossRef]
  35. Litasov, K.D.; Shatskiy, A.; Fei, Y.W.; Suzuki, A.; Ohtani, E.; Funakoshi, K. Pressure-Volume-Temperature equation of state of tungsten carbide to 32 GPa and 1673 K. J. Appl. Phys. 2010, 108, 053513. [Google Scholar] [CrossRef]
  36. Molina-Sánchez, A.; Wirtz, L. Phonons in single-layer and few-layer MoS2 and WS2. Phys. Rev. B 2011, 84, 155413. [Google Scholar] [CrossRef] [Green Version]
  37. Yan, Y.; Xia, B.; Qi, X.; Wang, H.; Xu, R.; Wang, J.-Y.; Zhang, H.; Wang, X. Nano-Tungsten carbide decorated graphene as co-catalysts for enhanced hydrogen evolution on molybdenum disulfide. Chem. Commun. 2013, 49, 4884–4886. [Google Scholar] [CrossRef] [PubMed]
  38. Cheng, Y.; Pang, K.; Wu, X.; Zhang, Z.; Xu, X.; Ren, J.; Huang, W.; Song, R. In Situ Hydrothermal Synthesis MoS2/Guar Gum Carbon Nanoflowers as Advanced Electrocatalysts for Electrocatalytic Hydrogen Evolution. ACS Sustain. Chem. Eng. 2018, 6, 8688–8696. [Google Scholar] [CrossRef]
  39. Shiva, K.; Ramakrishna Matte, H.S.S.; Rajendra, H.B.; Bhattacharyya, A.J.; Rao, C.N.R. Employing synergistic interactions between few-layer WS2 and reduced graphene oxide to improve lithium storage, cyclability and rate capability of Li-ion batteries. Nano Energy 2013, 2, 787–793. [Google Scholar] [CrossRef]
  40. Mak, K.F.; He, K.; Lee, C.; Lee, G.H.; Hone, J.; Heinz, T.F.; Shan, J. Tightly bound trions inmonolayer MoS2. Nat. Mater. 2013, 12, 207–211. [Google Scholar] [CrossRef]
  41. Zhang, C.; Ma, Y.; Zhang, X.; Abdolhosseinzadeh, S.; Sheng, H.; Lan, W.; Pakdel, A.; Heier, J.; Nüesch, F. Two-Dimensional Transition Metal Carbides and Nitrides (MXenes): Synthesis, Properties, and Electrochemical Energy Storage Applications. Energy Environ. Mater. 2020, 3, 29–55. [Google Scholar] [CrossRef]
  42. Moya, A.A. Identification of characteristic time constants in the initial dynamic response of electric double layer capacitors from high-frequency electrochemical impedance. J. Power Sour. 2018, 397, 124–133. [Google Scholar] [CrossRef]
  43. Samad, A.; Shafique, A.; Kim, H.J.; Shin, Y.-H. Superionic and electronic conductivity in monolayer W2C: Ab initio predictions. J. Mater. Chem. A 2017, 5, 11094–11099. [Google Scholar] [CrossRef]
Figure 1. SEM images of the (a) W2C/WS2_2h, (b) W2C/WS2_4h, (c) W2C/WS2_12h, and (d) WS2 NFs.
Figure 1. SEM images of the (a) W2C/WS2_2h, (b) W2C/WS2_4h, (c) W2C/WS2_12h, and (d) WS2 NFs.
Nanomaterials 10 01336 g001
Figure 2. (a) X-ray diffraction (XRD) patterns and (b) Raman spectra of the W2C/WS2 alloy nanoflowers (NFs) synthesized by heating at 250 °C for 2, 4, and 12 h.
Figure 2. (a) X-ray diffraction (XRD) patterns and (b) Raman spectra of the W2C/WS2 alloy nanoflowers (NFs) synthesized by heating at 250 °C for 2, 4, and 12 h.
Nanomaterials 10 01336 g002
Figure 3. (a) TEM image; (b,c) HRTEM image and (d) selected-area electron diffraction (SAED) pattern of the W2C/WS2_4h NF sample.
Figure 3. (a) TEM image; (b,c) HRTEM image and (d) selected-area electron diffraction (SAED) pattern of the W2C/WS2_4h NF sample.
Nanomaterials 10 01336 g003
Figure 4. XPS profiles of the W2C/WS2 _4h NF sample with (a) survey scan spectrum and high-resolution (b) W 4f, (c) S 2p, and (d) C 1s spectra.
Figure 4. XPS profiles of the W2C/WS2 _4h NF sample with (a) survey scan spectrum and high-resolution (b) W 4f, (c) S 2p, and (d) C 1s spectra.
Nanomaterials 10 01336 g004
Figure 5. Cyclic voltammograms of the (a) as-prepared WS2 NF, (b)W2C/WS2_2h NF, (c)W2C/WS2_4h NF, and (d) W2C/WS2_12h NF electrodes.
Figure 5. Cyclic voltammograms of the (a) as-prepared WS2 NF, (b)W2C/WS2_2h NF, (c)W2C/WS2_4h NF, and (d) W2C/WS2_12h NF electrodes.
Nanomaterials 10 01336 g005
Figure 6. (a) Initial voltage profiles and (b) cyclic performance of the as-prepared WS2 NF, W2C/WS2_2h NF, W2C/WS2_4h NF, and W2C/WS2_12h NF electrodes over 100 cycles.
Figure 6. (a) Initial voltage profiles and (b) cyclic performance of the as-prepared WS2 NF, W2C/WS2_2h NF, W2C/WS2_4h NF, and W2C/WS2_12h NF electrodes over 100 cycles.
Nanomaterials 10 01336 g006
Figure 7. Nyquist plots of the as-prepared WS2 NF, W2C/WS2_2h NF, W2C/WS2_4h NF, and W2C/WS2_12h NF electrodes with inset equivalent circuit.
Figure 7. Nyquist plots of the as-prepared WS2 NF, W2C/WS2_2h NF, W2C/WS2_4h NF, and W2C/WS2_12h NF electrodes with inset equivalent circuit.
Nanomaterials 10 01336 g007
Table 1. Comparison of resistance values extracted from equivalent circuit for the electrochemical impedance spectroscopic (EIS) measurement.
Table 1. Comparison of resistance values extracted from equivalent circuit for the electrochemical impedance spectroscopic (EIS) measurement.
SampleRsRctRSEI
WS2 NFs2.12627430
WC/WS2_2h3.04381367
WC/WS2_4h1.24312350
WC/WS2_12h5.41620422

Share and Cite

MDPI and ACS Style

Nguyen, T.P.; Kim, I.T. W2C/WS2 Alloy Nanoflowers as Anode Materials for Lithium-Ion Storage. Nanomaterials 2020, 10, 1336. https://doi.org/10.3390/nano10071336

AMA Style

Nguyen TP, Kim IT. W2C/WS2 Alloy Nanoflowers as Anode Materials for Lithium-Ion Storage. Nanomaterials. 2020; 10(7):1336. https://doi.org/10.3390/nano10071336

Chicago/Turabian Style

Nguyen, Thang Phan, and Il Tae Kim. 2020. "W2C/WS2 Alloy Nanoflowers as Anode Materials for Lithium-Ion Storage" Nanomaterials 10, no. 7: 1336. https://doi.org/10.3390/nano10071336

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