Next Article in Journal
Bending and Elastic Vibration of a Novel Functionally Graded Polymer Nanocomposite Beam Reinforced by Graphene Nanoplatelets
Next Article in Special Issue
Solid-State Solar Energy Conversion from WO3 Nano and Microstructures with Charge Transportation and Light-Scattering Characteristics
Previous Article in Journal
Ni-Based Catalyst Derived from NiAl Layered Double Hydroxide for Vapor Phase Catalytic Exchange between Hydrogen and Water
Previous Article in Special Issue
Homogeneous Core/Shell NiMoO4@NiMoO4 and Activated Carbon for High Performance Asymmetric Supercapacitor
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Facile Synthesis of Antimony Tungstate Nanosheets as Anodes for Lithium-Ion Batteries

1
Collaborative Innovation Center of Nonferrous Metals of Henan Province, Henan Key Laboratory of Non-Ferrous Materials Science & Processing Technology, School of Materials Science and Engineering, Henan University of Science and Technology, Luoyang 471023, China
2
National Joint Engineering Research Center for Abrasion Control and Molding of Metal Materials, Henan Key Laboratory of High-Temperature Structural and Functional Materials, Henan University of Science and Technology, Luoyang 471023, China
3
School of Agriculture, Henan University of Science and Technology, Luoyang 471023, China
4
School of Materials Science and Energy Engineering, Foshan University, Foshan 528000, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Nanomaterials 2019, 9(12), 1689; https://doi.org/10.3390/nano9121689
Submission received: 4 November 2019 / Revised: 20 November 2019 / Accepted: 20 November 2019 / Published: 25 November 2019
(This article belongs to the Special Issue Metal-Oxide Nanomaterials for Energy Application)

Abstract

:
Lithium-ion batteries (LIBs) have been widely used in the fields of smart phones, electric vehicles, and smart grids. With its opened Aurivillius structure, tungstate antimony oxide (Sb2WO6, SWO), constituted of {Sb2O2}2n+ and {WO4}2n−, is rarely investigated as an anode for lithium-ion batteries. In this work, Sb2WO6 with nanosheets morphology was successfully synthesized using a simple microwave hydrothermal method and systematically studied as an anode for lithium-ion batteries. The optimal SWO (SWO-60) exhibits a high specific discharge capacity and good rate capability. The good electrochemical performance could be ascribed to mesoporous nanosheets morphology, which is favorable for the penetration of the electrolyte and charge transportation. The results show that this nanostructured SWO is a promising anode material for LIBs.

1. Introduction

Lithium-ion batteries (LIBs) are widely used in the fields of portable devices, electric cars, and smart grids due to of their advantages, which include a higher working voltage and longer cycle life than many other energy storage systems [1,2,3,4,5,6,7,8,9]. However, the energy density of LIBs is still limited and may not meet demand, especially in the fields of high-performance electric vehicles and unmanned aerial vehicles, effectively hindering further development in these fields [10,11,12,13,14,15,16,17]. The reversible discharge capacity of commercial anodes (graphite), which is only ~372 mAh g−1, cannot meet the increasing demands for high-energy-density applications [18,19,20,21,22,23,24,25,26]. The use of proper anode materials with high capacities is a good strategy to solve the problem of LIBs [17]. Therefore, it is urgent that high-capacity negative materials for rechargeable lithium-ion batteries are developed and optimized.
As a potentially promising anode material for LIBs, transition metal oxides (TMO) have attracted broad attention due to their higher reversible capacity and proper redox voltage [25,26]. As one kind of TMO, oxysalts of tungsten, such as ZnWO4 [27], CaWO4 [28], FeWO4 [29], and Bi2WO6 [30,31], have been used as negative materials for lithium-ion batteries. For example, Zhang and colleagues [30] investigated Bi2WO6/graphene composite materials, which can exhibit 715.2 mAh g−1 at the first discharge at 50 mA g−1 and show excellent cycling performance. As a kind of tungstate, Sb2WO6 has a wide application in the field of photocatalysis [32,33,34] because of its unique layered structure and physicochemical properties [35,36]. Additionally, Sb2WO6 was recently reported as an anode for sodium-ion batteries with superior cycling and rate performance, delivering about 350 mA h g−1 after 100 cycles at 0.2 A g−1 [35]. Furthermore, the microwave hydrothermal method is well-known for effectively reducing reaction time and saving energy [37]. However, to the best of our knowledge, Sb2WO6 is rarely systematically investigated as an anode material for LIBs.
Herein, Sb2WO6 nanosheets were successfully fabricated through a microwave hydrothermal method, and their electrochemical performance was systematically studied as a negative electrode for LIBs. When used as a negative electrode material for LIBs, the optimal as-synthesized tungstate antimony oxide (SWO) sample exhibits a high reversible specific discharge capacity of ~698.95 mAh g−1 and remains at 424.22 mAh g−1 after 100 cycles at 0.2 A g−1. Furthermore, it exhibits a good rate performance.

2. Materials and Methods

2.1. Fabrication of Sb2WO6

All chemicals were directly used after purchase without further purification. As per the typical method, 0.1045 g Na2WO4·2H2O and 0.1446 g SbCl3 were separately dissolved in 20 mL and 10 mL of distilled water, respectively, and then stirred for 15 min. The Na2WO4·2H2O solution was added to the SbCl3 solution until a yellow suspension was formed, after which it was stirred for another 30 min. Next, the solution was added to a microwaveable water kettle (50 mL) and fixed in a microwave hydrothermal synthesizer (XH800, Beijing, China). A continuous microwave heating mode was set up to provide a stable heat source at a microwave power of 500 W. The temperature of the mixed solution quickly rose from room temperature to 160 °C and was maintained at 160 °C for different lengths of time (45 min, 60 min, and 75 min, which were denoted as SWO-45, SWO-60, and SWO-75, respectively). Afterwards, the resulting precipitates were filtered and washed several times with absolute ethanol and deionized water, and then the collected precipitation was dried overnight by lyophilization (LGJ-12, Beijing, China).

2.2. Material Characterization

Scanning electron micrographs were collected on a field-emission scanning electron microscope (FESEM, JSM-5610LV, JEOL, Akishima, Japan), and transmission electron micrographs and selected area electron diffraction patterns (SAED) were collected on a transmission electron microscope (TEM, JSM-2100F, 200 kV, Hitachinaka, Naka, Japan) to visually observe morphologies, particle sizes, etc. The crystallographic phase of the as-synthesized samples was characterized by X-ray diffraction (XRD, Bruker D8 ADVANCE, Cu kα source), and the 2θ-angle ranged from 10°–80°. Nitrogen adsorption-desorption isotherms were acquired by the Brunauer-Emmett-Teller (BET) method at 77 K from a Quadrasorb SI analyzer, and pore size distribution originated from the desorption branch based on the Barrett–Joyner–Halenda (BJH) theory.

2.3. Electrochemical Measurements

The electrochemical lithium storage performances were evaluated through a two-electrode system. Specifically, high-purity lithium foil (15.8 mm × 0.5 mm) was employed as the counter and reference electrode. After dissolving carboxymethylcellulose sodium (binder, CMC) in deionized water and stirring gently for 24 h to fabricate the working electrode, the slurry was made by mixing the Sb2WO6 samples, carbon black (Super P), and CMC at a weight ratio of 70:20:10. After fully stirring, the steady slurry was coated onto conductive copper foil and dried at 80 °C for 5 h. The dried electrodes were cut into a 12 mm-diameter disk to serve as a working electrode, and the active material loading in each disk was weighed to be 0.9–1.1 mg, corresponding to 0.8–0.97 mg cm−2. The batteries were assembled in the glove box with argon atmosphere. During assembly, the Celgard 2400 was used as a separator, while the electrolyte was made by dissolving 1M LiPF6 into a solvent consisting of ethylene carbonate (EC) and dimethyl carbonate (DMC) (EC/DMC, 1:1 in volume ratio). The galvanostatic discharge–charge measurements were performed on a LAND CT2001A in the voltage window from 0.01 to 3.0 V (vs. Li/Li+). An electrochemical impedance spectroscopy (EIS) was conducted ranging from 100 kHz to 0.01 Hz, and the cyclic voltammetry (CV) was measured in the range of 0.01–3.0 V (vs. Li/Li+) at 0.5 mV s−1. The two electrochemical tests discussed above were carried out on an electrochemistry workstation (CHI660E, Shanghai, China).

3. Results and Discussion

3.1. Structure and Morphology

Figure 1a illustrates the XRD patterns of the Sb2WO6 samples synthesized by a microwave hydrothermal method at different microwave heating times (45 min, 60 min, and 75 min). The peaks at 26.9° (211), 29.2° (002), 32.9° (310), and 36.6° (112) agree well with the diffraction pattern of Sb2WO6 (JCPDS card No.50-1553) [29]. Furthermore, no additional phase could be found, indicating that pure Sb2WO6 was successfully synthesized.
The Sb2WO6 samples synthesized by the different microwave heating times (SWO-45, SWO-60, and SWO-75) were also characterized by a nitrogen adsorption-desorption isotherms measurement. As depicted in Figure 1b–d, typical IV-type isotherms were obtained, indicating that almost no micropores existed in the Sb2WO6 samples [36]. A large number of mesopores (whose diameters are usually smaller than 30 nm) were observed in the pore-size distribution curves (inset in Figure 1b–d). The specific surface areas, mean pore sizes, and mean crystallite sizes of the three different SWO samples are listed in Table 1. According to the BET equation and the Barrett–Joyner–Halenda plot, the specific surface area of the SWO-60 sample was calculated as ~22.07 m2 g−1, and the mean pore size appeared at 3.9 nm. According to the Scherrer equation, the mean crystallite size was calculated as ~13.82 nm. In contrast, the specific surface area and the mean pore size of the SWO-45 sample was ~16.79 m2 g−1 and 3.7 nm, respectively. This specific surface area was much smaller than that of the SWO-60 sample, and the mean crystallite size of SWO-45 was higher than SWO-60. Although the specific surface area of the SWO-75 sample (~22.98 m2 g−1) was a little higher than the surface area of the SWO-60 sample, the main pore size was smaller than in the SWO-60 sample. It is believed that this mesoporous structure, with appropriate crystallite size, may be favorable for the transport and storage of lithium ions [38].
A scanning electron microscope and a transmission electron microscope were carried out to examine the microstructures and morphology of the SWO samples. As depicted in Figure 2a,b, the as-synthesized SWO-45 and SWO-60 samples exhibit similar nanosheet-like morphology. The nanosheets in SWO-45 (Figure 2a) are severely agglomerated and unevenly distributed, whereas the SWO-60 nanosheets are uniformly grown, and the thickness of the nanosheets is 12~21 nm, which might favor the charge transport and ion diffusion as well as the penetration of the electrolyte. In contrast, the SWO-75 sample (Figure 2c) seems to be of nanoparticle assembly, the reason for which may be that the microwave reaction time was too long, resulting in the overgrowth of nanosheets. As depicted in Figure 2d, the transmission electron microscope image further confirms flake morphology of the SWO-60 sample. A high-resolution transmission electron microscope (HRTEM) image is depicted in Figure 2e, with interplanar spacing of 3.3 Å in a single nanosheet, which corresponds to the spacing between the (211) crystal plane of the Sb2WO6. Figure 2f shows the SAED of SWO-60, indicating that the crystalline structure of SWO-60 is polycrystalline, which agrees well with the X-ray diffraction patterns depicted in Figure 1a.
To further investigate the chemical composition and valence state of SWO-60, an X-ray photoelectron spectroscopy (XPS) was conducted. In Figure 3b, the binding energies of 530.9 eV and 540.3 eV are in accordance with the Sb 3d5/2 and Sb 3d3/2 with 3+ oxidation state, respectively [39]. In addition, the peak at 531.9 eV is the characteristic of O 1s. In Figure 3c, the peak at 767.9 eV could be attributed to the Sb 3p3/2 binding energy of Sb2WO6. Additionally, the peaks at 35.9 eV, 38.1 eV, and 34.5 eV are characteristics of W 4f7/2, W 4f5/2, and Sb 4d, respectively, and W 4f7/2 and W 4f5/2 are in accordance with the oxidation state of W6+ [40,41].

3.2. Electrochemical Performance

Electrochemical tests were conducted to examine the performances of different SWO samples in lithium-ion batteries (Figure 4). The cyclic voltammetry (CV) of a SWO-60 electrode at a scan rate of 0.5 mV s−1 between 0.01 V and 3 V is depicted in Figure 4a. The reduction peak located at ~1.125 V can be attributed to the formation of solid electrolyte interface (SEI) film, during which Sb2WO6 was decomposed and antimony oxide was converted into metallic antimony [35]. The reduction peak centered at ~0.43V could be attributed to the alloying-reaction that took place between Li and Sb, and the oxidation peak of 1.254V is related to the dealloying of Li3Sb [35,42]. As we can see, from the third cycle on, the cyclic voltammetry curves are well overlapped, indicating good reversibility of the SWO-60 electrode.
The discharge–charge profile of the SWO-60 electrode in the 1st, 2nd, 5th, 50th, and 100th cycles at 0.2 A g−1 is shown in Figure 4b. During the first cycle, the initial specific discharge and charge capacities of the Sb2WO6 electrode were ~875.06 and ~678.42 mAh g−1, respectively, corresponding to 77.52% in initial coulombic efficiency, which could be attributed to the formation of a solid electrolyte interphase during the first discharge process. As shown in Figure 4b, there are two well-defined plateaus at ~1.3 and ~0.7 V in the first discharge curve. The plateau of ~1.3 V disappears after the first cycle, while the ~0.7 V plateau remains in the following cycles, which is consistent with the CV discussed above. The cycling stabilities of different Sb2WO6 electrodes are shown in Figure 4c. The initial specific discharge capacities of the SWO-45, SWO-60, and SWO-75 electrodes are 246.03, 875.06, and 764.5 mAh g−1 at 0.2 A g−1, respectively. The discharge capacities of the SWO-45 and SWO-75 electrodes decrease rapidly, the specific discharge capacities of which undergo a decrease of ~49% and ~69%, respectively, from the 2nd to the 100th cycle, that is, they decrease at a rate of ~1.9 and ~4.4 mAh g−1 per cycle, respectively. In contrast, the reversible capacity of the SWO-60 electrode still delivers 424.22 mAh g−1 at 0.2 A g−1 after 100 cycles, corresponding to a decrease of ~39%, which is less than the other two electrodes, which may be associated with the uniform distributed nanosheets structure. However, as mentioned in the literature, the cycling stability of SWO-60 is inferior to Sb2WO6 for sodium-ion batteries [29], which may be attributed to the poor conductivity of SWO-60 [27], and its cycling stability could be improved through compositing with amorphous carbon, graphene, and other materials with higher conductivity, such as conducting polymers. Figure 4d illustrates the rate-performance of three electrodes at current densities ranging from 0.1 to 1.1 A g−1. The SWO-60 electrode displays reversible discharge capacities of 743.9, 513.7, and 462.2 mAh g−1 at 0.1, 0.7, and 1.1 A g−1, respectively. The SWO-60 electrode can still deliver 703.6 mA h g−1 when the current density reverts to 0.1 A g−1, which is much higher than the SWO-45 and SWO-75 electrodes. The excellent rate performance of SWO-60 could be associated with its unique nanosheets morphology, which facilitates rapid lithium-ion conduction and transportation [43].
In order to investigate the electrochemical kinetics of different kinds of electrodes, electrochemical impedance spectroscopy measurements were carried out from 0.1 Hz to 100 kHz in frequency. Figure 5a,b shows the Nyquist plots of three kinds of electrodes in initial their states and after 200 discharge–charge cycling at 0.2 A g−1, respectively. In the impedance spectra, the semicircles in the high-frequency region correspond to the charge-transfer resistance (Rct) in the interface of electrodes and electrolyte, and the inclined lines during the low-frequency region are attributed to insertion of lithium ions into the negative electrode [44]. In Figure 5a, the charge transfer resistance of a battery with a SWO-60 electrode (332.8 Ω) is much smaller than those of the SWO-45 (998.4 Ω) and SWO-75 (477.3 Ω) batteries, which is consistent with the good cycling stability of the SWO-60 electrode. Figure 5b shows the EIS of SWO-45, SWO-60, and SWO-75 cells after 200 cycles at 0.2 A g−1. After cycling, the cell with the SWO-60 electrode also exhibited the lowest charge transfer resistance, which is in good agreement with the fact that the SWO-60 electrode showed the best electrochemical properties.
To investigate the effect of different microwave hydrothermal times on the structural stability of electrodes, the morphology of electrodes was examined after 200 cycles at 0.2 A g−1 (Figure 6). Figure 6a–c shows the morphology of the SWO-45, SWO-60, and SWO-75 electrodes after cycling, respectively. After 200 cycles, it was found that there were many large cracks and pulverizations in the SWO-45 and SWO-75 electrodes, as shown in Figure 6a,c, respectively. By contrast, the SWO-60 electrode retained its structural integrity even after 200 cycles, as shown in Figure 6b.

4. Conclusions

In summary, mesoporous Sb2WO6 nanosheets were successfully fabricated using a microwave hydrothermal method and were systematically investigated as negative electrodes for LIBs. When the microwave reaction time is 60 min, SWO-60 exhibits the highest reversible capacity and the best rate performance. The good electrochemical performance of the SWO-60 electrode may be ascribed to its uniformly distributed nanosheets structure. Due to the limited inherent electronic conductivity of SWO, our future work will focus on combining SWO with other materials with higher conductivities, such as heteroatom-doped nanocarbon materials and graphene, to further improve their electrochemical performance.

Author Contributions

Y.L., W.Z., and F.R. conceived and designed the experiments; Y.W., F.W., and Z.L. performed the experiments; Y.W., J.L., and K.P. analyzed the data; Y.W. wrote the paper; Y.L., G.W., M.C., and S.W. revised the paper. All the authors have discussed the results and commented on the manuscript.

Funding

This work was supported by the Program for Changjiang Scholars and Innovative Research Team in University (IRT_16R21), the Chinese 02 Special Fund (2017ZX02408003), the Key Research Projects of Colleges and Universities in the Henan Province (15A430023), the Scientific and Technological Project of Henan Province (182102210297), Open Fund of National Joint Engineering Research Center for abrasion control and molding of metal materials (HKDNM201807), Scientific Research Starting Foundation for Ph.D. of Henan University of Science and Technology (13480065), Science Foundation for Youths of Henan University of Science and Technology (2013QN006), and the Student Research Training Plan of Henan University of Science and Technology (2018029, 2019031).

Acknowledgments

Y. W. is grateful for the valuable discussion and help from Huijie Wei from School of Materials Science and Engineering, Henan University of Science and Technology.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chu, S.; Cui, Y.; Liu, N. The path towards sustainable energy. Nat. Mater. 2017, 16, 16–22. [Google Scholar] [CrossRef]
  2. Tarascon, J.M.; Armand, M. Issues and challenges facing rechargeable lithium batteries. Nature 2001, 414, 359–367. [Google Scholar] [CrossRef] [PubMed]
  3. Armand, M.; Tarascon, J.M. Building better batteries. Nature 2008, 451, 652–657. [Google Scholar] [CrossRef] [PubMed]
  4. Goodenough, J.B.; Kim, Y. Challenges for Rechargeable Li Batteries. Chem. Mater. 2010, 22, 587–603. [Google Scholar] [CrossRef]
  5. Wu, H.B.; Chen, J.S.; Hng, H.H.; Lou, X.W. Nanostructured metal oxide-based materials as advanced anodes for lithium-ion batteries. Nanoscale 2012, 4, 2526–2542. [Google Scholar] [CrossRef] [PubMed]
  6. Cheng, C.; Geng, T.; Du, C.; Zuo, P.; Cheng, X.; Ma, Y.; Yin, G. Oxygen vacancies in SnO2 surface coating to enhance the activation of layered Li-Rich Li1.2 Mn0.54 Ni0.13 Co0.13 O2 cathode material for Li-ion batteries. J. Power Sources 2016, 331, 91–99. [Google Scholar] [CrossRef]
  7. Wang, Y.; Wang, Z.; Lei, D.; Lv, W.; Zhao, Q.; Ni, B.; Liu, Y.; Li, B.; Kang, F.; He, Y.-B. Spherical Li Deposited inside 3D Cu Skeleton as Anode with Ultrastable Performance. ACS Appl. Mater. Interfaces 2018, 10, 20244–20249. [Google Scholar] [CrossRef]
  8. Zheng, J.-C.; Yang, Z.; He, Z.-J.; Tong, H.; Yu, W.-J.; Zhang, J.-F. In situ formed LiNi0.8Co0.15Al0.05O2@Li4SiO4 composite cathode material with high rate capability and long cycling stability for lithium-ion batteries. Nano Energy 2018, 53, 613–621. [Google Scholar] [CrossRef]
  9. Liu, Y.; Zhai, X.; Yang, K.; Wang, F.; Wei, H.; Zhang, W.; Ren, F.; Pang, H. Mesoporous NH4NiPO4 center dot H2O for High-Performance Flexible All-Solid-State Asymmetric Supercapacitors. Front. Chem. 2019, 7. [Google Scholar] [CrossRef]
  10. Guo, X.T.; Zhang, Y.Z.; Zhang, F.; Li, Q.; Anjum, D.H.; Liang, H.F.; Liu, Y.; Liu, C.S.; Alshareef, H.N.; Pang, H. A novel strategy for the synthesis of highly stable ternary SiOx composites for Li-ion-battery anodes. J. Mater. Chem. A 2019, 7, 15969–15974. [Google Scholar] [CrossRef]
  11. Peng, T.; Liu, C.; Hou, X.Y.; Zhang, Z.W.; Wang, C.L.; Yan, H.L.; Lu, Y.; Liu, X.M.; Luo, Y.S. Control Growth of Mesoporous Nickel Tungstate Nanofiber and Its Application as Anode Material for Lithium-Ion Batteries. Electrochim. Acta 2017, 224, 460–467. [Google Scholar] [CrossRef]
  12. Hao, X.G.; Zhao, Q.; Su, S.M.; Zhang, S.Q.; Ma, J.B.; Shen, L.; Yu, Q.P.; Zhao, L.; Liu, Y.; Kang, F.Y.; et al. Constructing Multifunctional Interphase between Li1.4Al0.4Ti1.6(PO4)3 and Li Metal by Magnetron Sputtering for Highly Stable Solid-State Lithium Metal Batteries. Adv. Energy Mater. 2019, 9, 1901604. [Google Scholar] [CrossRef]
  13. Peng, T.; Hou, X.Y.; Liu, C.; Yu, Q.H.; Luo, R.J.; Yan, H.L.; Lu, Y.; Liu, X.M.; Luo, Y.S. Controlled synthesis of hierarchical CoMn2O4 nanostructures for flexible all-solid-state battery-type electrodes. J. Solid State Electrochem. 2017, 21, 1579–1587. [Google Scholar] [CrossRef]
  14. Wu, N.T.; Shen, J.K.; Sun, L.; Yuan, M.Y.; Shao, Y.Y.; Ma, J.M.; Liu, G.L.; Guo, D.L.; Liu, X.M.; He, Y.B. Hierarchical N-doped graphene coated 1D cobalt oxide microrods for robust and fast lithium storage at elevated temperature. Electrochim. Acta 2019, 310, 70–77. [Google Scholar] [CrossRef]
  15. Kang, B.; Ceder, G. Battery materials for ultrafast charging and discharging. Nature 2009, 458, 190–193. [Google Scholar] [CrossRef]
  16. Wu, H.; Chan, G.; Choi, J.W.; Ryu, I.; Yao, Y.; McDowell, M.T.; Lee, S.W.; Jackson, A.; Yang, Y.; Hu, L.B.; et al. Stable cycling of double-walled silicon nanotube battery anodes through solid-electrolyte interphase control. Nat. Nanotechnol. 2012, 7, 309–314. [Google Scholar] [CrossRef]
  17. Whittingham, M.S. Lithium batteries and cathode materials. Chem. Rev. 2004, 104, 4271–4301. [Google Scholar] [CrossRef]
  18. Kwasi-Effah, C.C.; Rabczuk, T. Dimensional analysis and modelling of energy density of lithium-ion battery. J. Energy Storage 2018, 18, 308–315. [Google Scholar] [CrossRef]
  19. Liu, Y.; Wei, H.; Wang, C.; Wang, F.; Wang, H.; Zhang, W.; Wang, X.; Yan, C.; Kim, B.H.; Ren, F. Nitrogen-Doped Carbon Coated WS2 Nanosheets as Anode for High-Performance Sodium-Ion Batteries. Front. Chem. 2018, 6. [Google Scholar] [CrossRef]
  20. Wang, F.; Liu, Y.; Zhao, Y.; Wang, Y.; Wang, Z.; Zhang, W.; Ren, F. Facile Synthesis of Two-Dimensional Porous MgCo2O4 Nanosheets as Anode for Lithium-Ion Batteries. Appl. Sci. 2018, 8, 22. [Google Scholar] [CrossRef]
  21. Liu, G.; Cui, J.; Luo, R.; Liu, Y.; Huang, X.; Wu, N.; Jin, X.; Chen, H.; Tang, S.; Kim, J.-K.; et al. 2D MoS2 grown on biomass-based hollow carbon fibers for energy storage. Appl. Surf. Sci. 2019, 469, 854–863. [Google Scholar] [CrossRef]
  22. Liu, Y.; Wang, H.C.; Yang, K.K.; Yang, Y.N.; Ma, J.Q.; Pan, K.M.; Wang, G.X.; Ren, F.Z.; Pang, H. Enhanced Electrochemical Performance of Sb2O3 as an Anode for Lithium-Ion Batteries by a Stable Cross-Linked Binder. Appl. Sci. 2019, 9, 2677. [Google Scholar] [CrossRef]
  23. Kasnatscheew, J.; Wagner, R.; Winter, M.; Cekic-Laskovic, I. Interfaces and Materials in Lithium Ion Batteries: Challenges for Theoretical Electrochemistry. Top. Curr. Chem. 2018, 376, 29. [Google Scholar] [CrossRef]
  24. Wang, H.; Wei, Y.J.; Wang, J.T.; Long, D.H. Polymer-chelation synthesis of compositionally homogeneous LiNi1/3Co1/3Mn1/3O2 crystals for lithium-ion cathode. Electrochim. Acta 2018, 269, 724–732. [Google Scholar] [CrossRef]
  25. Wang, H.; Yang, X.; Wu, Q.; Zhang, Q.; Chen, H.; Jing, H.; Wang, J.; Mi, S.-B.; Rogach, A.L.; Niu, C. Encapsulating Silica/Antimony into Porous Electrospun Carbon Nanofibers with Robust Structure Stability for High-Efficiency Lithium Storage. ACS Nano 2018, 12, 3406–3416. [Google Scholar] [CrossRef]
  26. Zeng, C.; Weng, W.; Lv, T.; Xiao, W. Low-Temperature Assembly of Ultrathin Amorphous MnO2 Nanosheets over Fe2O3 Spindles for Enhanced Lithium Storage. ACS Appl. Mater. Interfaces 2018, 10, 30470–30478. [Google Scholar] [CrossRef]
  27. Shim, H.W.; Cho, I.S.; Hong, K.S.; Lim, A.H.; Kim, D.W. Wolframite-type ZnWO4 Nanorods as New Anodes for Li-Ion Batteries. J. Phys. Chem. C 2011, 115, 16228–16233. [Google Scholar] [CrossRef]
  28. Sharma, N.; Rao, G.V.S.; Chowdari, B.V.R. Electrochemical properties of carbon-coated CaWO4 versus Li. Electrochim. Acta 2005, 50, 5305–5312. [Google Scholar] [CrossRef]
  29. Hyun-Woo, S.; In-Sun, C.; Kug Sun, H.; Won Il, C.; Dong-Wan, K. Li electroactivity of iron (II) tungstate nanorods. Nanotechnology 2010, 21, 465602. [Google Scholar]
  30. Zhang, L.S.; Bai, Q.L.; Jin, K.; Wang, L.Z.; Zhang, Y.; Song, Y.H. Synthesis and electrochemical performance of Bi2WO6/graphene composite as anode material for lithium-ion batteries. Mater. Lett. 2015, 141, 88–91. [Google Scholar] [CrossRef]
  31. Zhang, L.S.; Zhao, Z.J.; Xia, T.C.; Zhang, S.L.; Li, X.F.; Zhang, A.Q. Anchoring Bi2WO6 nanoparticles on 3D graphene frameworks for enhanced lithium storage. Mater. Lett. 2018, 210, 345–349. [Google Scholar] [CrossRef]
  32. Singh, A.; Dutta, D.P.; Roy, M.; Tyagi, A.K.; Fulekar, M.H. Sonochemical synthesis, characterization, and photocatalytic properties of Bi2−xSbxWO6 nanorods. J. Mater. Sci. 2014, 49, 2085–2097. [Google Scholar] [CrossRef]
  33. Yang, C.Y.; Yang, X.J.; Li, F.; Li, T.H.; Cao, W. Controlled synthesis of hierarchical flower-like Sb2WO6 microspheres: Photocatalytic and superhydrophobic property. J. Ind. Eng. Chem. 2016, 39, 93–100. [Google Scholar] [CrossRef]
  34. Hu, S.P.; Xu, C.Y.; Ma, F.X.; Cao, L.; Zhen, L. Solvothermal synthesis of orthorhombic Sb2WO6 hierarchical structures and their visible-light-driven photocatalytic activity. Dalton Trans. 2014, 43, 8439–8445. [Google Scholar] [CrossRef]
  35. Wang, P.; Xie, S.M.; She, Y.Y.; Fan, W.G.; Leung, M.K.H.; Wane, H.K. Microwave-Hydrothermal Synthesis of Hierarchical Sb2WO6 Nanostructures as a New Anode Material for Sodium Storage. Chemistryselect 2019, 4, 1078–1083. [Google Scholar] [CrossRef]
  36. Chen, S.Q.; Zhou, M.Y.; Li, T.H.; Cao, W. Synthesis of Ag-loaded Sb2WO6 microsphere with enhanced photocatalytic ability for organic dyes degradations under different light irradiations. J. Mol. Liq. 2018, 272, 27–36. [Google Scholar] [CrossRef]
  37. Liu, Y.; Hori, A.; Kusaka, S.; Hosono, N.; Li, M.; Guo, A.; Du, D.; Li, Y.; Yang, W.; Ma, Y.; et al. Microwave-Assisted Hydrothermal Synthesis of Al(OH)(1,4-NDC) Membranes with Superior Separation Performances. Chem. Asian J. 2019. [Google Scholar] [CrossRef]
  38. Park, H.C.; Lee, K.H.; Lee, Y.W.; Kim, S.J.; Kim, D.M.; Kim, M.C.; Park, K.W. Mesoporous molybdenum nitride nanobelts as an anode with improved electrochemical properties in lithium ion batteries. J. Power Sources 2014, 269, 534–541. [Google Scholar] [CrossRef]
  39. Lu, X.; Wang, Z.Y.; Lu, L.; Yang, G.; Niu, C.M.; Wang, H.K. Synthesis of Hierarchical Sb2MoO6 Architectures and Their Electrochemical Behaviors as Anode Materials for Li-Ion Batteries. Inorg. Chem. 2016, 55, 7012–7019. [Google Scholar] [CrossRef]
  40. Wu, X.; Yao, S.Y. Flexible electrode materials based on WO3 nanotube bundles for high performance energy storage devices. Nano Energy 2017, 42, 143–150. [Google Scholar] [CrossRef]
  41. Su, Y.B.; Zhang, H.J.; Liang, P.; Liu, K.; Cai, M.Y.; Huang, Z.Y.; Wang, C.A.; Zhong, M.L. A new binder-free and conductive-additive-free TiO2/WO3-W integrative anode material produced by laser ablation. J. Power Sources 2018, 378, 362–368. [Google Scholar] [CrossRef]
  42. Lu, X.; Wang, P.; Liu, K.; Niu, C.M.; Wang, H.K. Encapsulating nanoparticulate Sb/MoOx into porous carbon nanofibers via electrospinning for efficient lithium storage. Chem. Eng. J. 2018, 336, 701–709. [Google Scholar] [CrossRef]
  43. Bai, J.; Zhao, B.C.; Zhou, J.F.; Fang, Z.T.; Li, K.Z.; Ma, H.Y.; Dai, J.M.; Zhu, X.B.; Sun, Y.P. Improved Electrochemical Performance of Ultrathin MoS2 Nanosheet/Co Composites for Lithium-Ion Battery Anodes. Chemelectrochem 2019, 6, 1930–1938. [Google Scholar] [CrossRef]
  44. Jin, J.; Huang, S.-Z.; Shu, J.; Wang, H.-E.; Li, Y.; Yu, Y.; Chen, L.-H.; Wang, B.-J.; Su, B.-L. Highly porous TiO2 hollow microspheres constructed by radially oriented nanorods chains for high capacity, high rate and long cycle capability lithium battery. Nano Energy 2015, 16, 339–349. [Google Scholar] [CrossRef]
Figure 1. (a) X-ray diffractometry pattern of as-synthesized tungstate antimony oxide (SWO)-45, SWO-60, and SWO-75 samples. Nitrogen adsorption-desorption isotherms and pore-size distribution (inset) of (b) SWO-45, (c) SWO-60, and (d) SWO-75 samples.
Figure 1. (a) X-ray diffractometry pattern of as-synthesized tungstate antimony oxide (SWO)-45, SWO-60, and SWO-75 samples. Nitrogen adsorption-desorption isotherms and pore-size distribution (inset) of (b) SWO-45, (c) SWO-60, and (d) SWO-75 samples.
Nanomaterials 09 01689 g001
Figure 2. (ac) Scanning electron microscope (SEM) images of as-prepared SWO-45, SWO-60, and SWO-75 samples; (d) transmission electron microscope (TEM) image of SWO-60; (e) high-resolution transmission electron microscope (HRTEM) image; and (f) selected area electron diffraction (SAED) pattern of the SWO-60 sample.
Figure 2. (ac) Scanning electron microscope (SEM) images of as-prepared SWO-45, SWO-60, and SWO-75 samples; (d) transmission electron microscope (TEM) image of SWO-60; (e) high-resolution transmission electron microscope (HRTEM) image; and (f) selected area electron diffraction (SAED) pattern of the SWO-60 sample.
Nanomaterials 09 01689 g002
Figure 3. (a) X-ray photoelectron spectroscopy (XPS) survey of the SWO-60 sample, and the high-resolution XPS spectra of (b) Sb 3d and O 1s, (c) Sb 3p, and (d) Sb 4d and W 4f.
Figure 3. (a) X-ray photoelectron spectroscopy (XPS) survey of the SWO-60 sample, and the high-resolution XPS spectra of (b) Sb 3d and O 1s, (c) Sb 3p, and (d) Sb 4d and W 4f.
Nanomaterials 09 01689 g003
Figure 4. (a) Cyclic voltammetry curves of the SWO-60 electrode at 0.5 mV s−1; (b) cdischarge–charge profile of SWO-60 electrode at different cycles at 0.2 A g−1; (c) cycling performances of SWO-45, SWO-60, and SWO-60 electrodes at 0.2 A g−1; and (d) rate performance of SWO-45, SWO-60, and SWO-60 electrodes at current density ranging from 0.1–1.1 A g−1.
Figure 4. (a) Cyclic voltammetry curves of the SWO-60 electrode at 0.5 mV s−1; (b) cdischarge–charge profile of SWO-60 electrode at different cycles at 0.2 A g−1; (c) cycling performances of SWO-45, SWO-60, and SWO-60 electrodes at 0.2 A g−1; and (d) rate performance of SWO-45, SWO-60, and SWO-60 electrodes at current density ranging from 0.1–1.1 A g−1.
Nanomaterials 09 01689 g004
Figure 5. Nyquist plots of different electrodes (a) before cycling and (b) after 200 cycles at 0.2 A g−1.
Figure 5. Nyquist plots of different electrodes (a) before cycling and (b) after 200 cycles at 0.2 A g−1.
Nanomaterials 09 01689 g005
Figure 6. Scanning electron microscope images of (a) SWO-45, (b) SWO-60, and (c) SWO-75 electrodes after 200 cycles at 0.2 A g−1.
Figure 6. Scanning electron microscope images of (a) SWO-45, (b) SWO-60, and (c) SWO-75 electrodes after 200 cycles at 0.2 A g−1.
Nanomaterials 09 01689 g006
Table 1. Comparison of Sb2WO6 samples synthesized by different microwave heating times in terms of specific surface area, mean pore size, and mean crystallite size.
Table 1. Comparison of Sb2WO6 samples synthesized by different microwave heating times in terms of specific surface area, mean pore size, and mean crystallite size.
MaterialsSpecific Surface Area (m2/g)Mean Pore Size (nm)Mean Crystallite Size (nm)
SWO-4516.793.715.37
SWO-6022.073.913.82
SWO-7522.983.613.16

Share and Cite

MDPI and ACS Style

Liu, Y.; Wang, Y.; Wang, F.; Lei, Z.; Zhang, W.; Pan, K.; Liu, J.; Chen, M.; Wang, G.; Ren, F.; et al. Facile Synthesis of Antimony Tungstate Nanosheets as Anodes for Lithium-Ion Batteries. Nanomaterials 2019, 9, 1689. https://doi.org/10.3390/nano9121689

AMA Style

Liu Y, Wang Y, Wang F, Lei Z, Zhang W, Pan K, Liu J, Chen M, Wang G, Ren F, et al. Facile Synthesis of Antimony Tungstate Nanosheets as Anodes for Lithium-Ion Batteries. Nanomaterials. 2019; 9(12):1689. https://doi.org/10.3390/nano9121689

Chicago/Turabian Style

Liu, Yong, Yue Wang, Fei Wang, Zhenxiao Lei, Wanhong Zhang, Kunming Pan, Jing Liu, Min Chen, Guangxin Wang, Fengzhang Ren, and et al. 2019. "Facile Synthesis of Antimony Tungstate Nanosheets as Anodes for Lithium-Ion Batteries" Nanomaterials 9, no. 12: 1689. https://doi.org/10.3390/nano9121689

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