Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 + 1777 word(s) 1777 2022-02-21 09:53:37 |
2 format correct -24 word(s) 1753 2022-02-22 04:11:24 | |
3 Roll back entry to make some changes. + 133 word(s) 1910 2022-02-22 04:13:23 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Zhao, T. Sea Urchin-like Si@MnO2@rGO. Encyclopedia. Available online: https://encyclopedia.pub/entry/19686 (accessed on 03 July 2024).
Zhao T. Sea Urchin-like Si@MnO2@rGO. Encyclopedia. Available at: https://encyclopedia.pub/entry/19686. Accessed July 03, 2024.
Zhao, Tianming. "Sea Urchin-like Si@MnO2@rGO" Encyclopedia, https://encyclopedia.pub/entry/19686 (accessed July 03, 2024).
Zhao, T. (2022, February 21). Sea Urchin-like Si@MnO2@rGO. In Encyclopedia. https://encyclopedia.pub/entry/19686
Zhao, Tianming. "Sea Urchin-like Si@MnO2@rGO." Encyclopedia. Web. 21 February, 2022.
Sea Urchin-like Si@MnO2@rGO
Edit

An unique structure which can effectively reduce the volume change of Si, extend the cycle life and increase the lithium-ion battery capacity.

Si lithium-ion battery

1. Anode design and morphology characterization

Figure 1 shows a schematic of the preparation of the sea urchin-like Si@MnO2@rGO composites. First, GO sheets were reduced to rGO by a hydrothermal method with NaBH4. Then, the rGO was collected and dispersed in deionized water (Solution A). Next, Si nanospheres were dispersed in a solution of water, alcohol, and the KH550 silane coupling agent (Solution B). The composites were well grafted by KH550, improving the conductivity.
Figure 1. Concise preparation schematics of the Si@MnO2@ reduced graphene oxide (rGO) composites.
Figure 2 shows the morphologies of the materials grown at three different temperatures. The morphology of Si@MnO2-20°C is shown in Figure 2a. The MnO2 formed disordered and entangled filaments on the surface of Si nanospheres. Figure 2b shows Si@MnO2-20°C at a high magnification. The surface of the material was rough and the diameter of Si@MnO2-20°C particles was 200 nm. Figure 2c is an SEM image of Si@MnO2@rGO-20°C, showing that the rGO covered the Si@MnO2-20°C. However, some MnO2 nanowires appeared on the surface of rGO, which suggested that the material was not stable. Figure 2d,e show the SEM images of Si@MnO2-50°C at different magnifications, respectively. As shown in Figure 2d, the MnO2 formed a sea urchin-like shell coating on the surface of Si nanospheres. Figure 2e shows the sea urchin-like Si@MnO2-50°C. The stings (MnO2) were 100–300 nm in length. Figure 2f shows Si@MnO2@rGO-50°C. The rGO wrapped the Si@MnO2-50°C without obvious damage to the structure. Interestingly, this rGO surface was clean, without any MnO2 nanowires, indicating the sea urchin-like structure was more stable. MnO2 nanowires were vertically oriented on the outer layer of the Si nanospheres, imbedding rGO layers, forming a stable structure, which can buffer the excessive expansion of the Si nanospheres during the charging/discharging process and extend the batteries’ lifetime. Figure 2g shows the SEM image of Si@MnO2-160°C. The MnO2 nanowires grew longer and coated the surface of the Si nanospheres. Figure 2h is the enlarged view of Figure 2g. The plentiful MnO2 nanowires were interleaved. Figure 2i shows the images of Si@MnO2@rGO-160°C. 
Figure 2. (a) SEM image of the Si@MnO2-20°C composite. (b) Enlarged view of the SEM image of the Si@MnO2-20°C composite. (c) SEM image of the Si@MnO2@rGO-20°C composite. (d) SEM image of the Si@MnO2-50°C composite. (e) Enlarged view of the SEM image of the Si@MnO2-50°C composite. (f) SEM image of the Si@MnO2@rGO-50°C composite. (g) SEM image of the Si@MnO2-160°C composite. (h) Enlarged view of the SEM image of the Si@MnO2-160°C composite. (i) SEM image of the Si@MnO2@rGO-160°C composite.
Figure 3a shows the Energy Dispersive Spectrometer (EDS) analysis of the ball-milled material, using aluminum foil as a substrate, and the inset shows the atomic percentages of elements in the materials. The inset shows that the Si@MnO2@rGO-50°C contained ~48% Si, ~21% rGO, and ~31% MnO2. Figure 3b,c show HRTEM images of the Si@MnO2@rGO-50°C composite at different magnifications. Si nanospheres were coated with rGO and MnO2. The lattice fringes corresponded to the (111) plane of Si, having a separation of 0.31 nm [1][2]. In addition, lattice fringes having an interplanar spacing of 0.47 nm can be seen, and this is consistent with the (200) plane of MnO2 [3][4][5]. The element mapping of the Si@MnO2@rGO-50°C composite is shown in Figure 3d–h. Mn element and O element contributed to MnO2; Si element contributed to Si nanospheres. C element dispersed throughout the image may have contributed to both the substrate and rGO.
Figure 3. (a) Energy Dispersive Spectrometer (EDS) spectra of Si@MnO2@rGO-50°C. Inset showing the respective substance of Si@MnO2@rGO-50°C. (b,c) HRTEM images of the Si@MnO2@rGO-50°C composite at different magnifications. (dh) Elemental mapping of the Si@MnO2@rGO-50°C composite.
Figure 4a shows the XRD patterns of rGO, Si, Si@MnO2-50°C, and Si@MnO2@rGO-50°C, respectively. In the XRD pattern of Si@MnO2-50°C, the pronounced peaks at 28.5°, 47.4°, 56.2°,58.9°, 69.2°, 76.5°, and 88.2° corresponded to the (111), (220), (311), (222), (400), (331), and (422) planes of Si (PDF#77-2108). In addition, the characteristic peaks at 12.7°, 18.0°, 28.7°, 37.6°, 41.1°, 49.8°, 59.5°, 65.5°, 68.5°, and 72.5° corresponded to the (110), (200), (310), (121), (420), (411), (260), (002), (202), and (631) planes of MnO2 (PDF#72-1982). The XRD pattern of the Si@MnO2@rGO-50°C composites contained peaks corresponding to Si and MnO2, as well as the broad peak ranges from 20° to 30°, which were indexed to the standard peaks of rGO [6][7][8][9]Figure 4b shows the XPS spectrum of the Si@MnO2@rGO-50°C composites, which revealed the presence of Si, Mn, O, and C, corresponding to element mapping (Figure 3d–g). Figure 4c,d shows the high-resolution XPS spectra. Figure 4c shows the Si 2p spectrum, whose peak at 99.7 eV was related to Si-Si bonds; two small peaks at 101.9 and 103.6 eV contributed to organic Si and Si-O, respectively, which may have been caused by the slight oxidation of Si in the thermal-treated process. Figure 4d shows the Mn 2p spectrum, which contained two spin–orbit peaks corresponding to Mn 2p3/2 (642.5 eV) and Mn 2p1/2 (654.1 eV) of MnO2, whose separation between these peaks was 11.6 eV.
Figure 4. (a) X-ray diffraction (XRD) patterns of the Si, Si@MnO2-50°C rGO, and Si@MnO2@rGO-50°C, respectively. (bd) XPS spectra of the Si@MnO2@rGO-50°C.

2. Lithium-Ion Battery Performance

A CV experiment carried out to further evaluate the lithium storage behavior has been shown in Figure 5. Figure 5a shows the CV curves of Si@MnO2@rGO-50°C as the independent anodes of lithium-ion batteries for the first four cycles at a scan rate of 0.1 mV s-1 between 0.1 V and 3.2 V. In the first cycle, a clear cathodic peak at 0.16 V corresponds to the lithium alloying process of crystalline Si and the formation of an amorphous LixSi phase, and a clear cathodic peak at 0.10 V corresponds to the formation of the Li15Si4 phase. An anodic peak at 0.24 V is related to the delithiation of Li15Si4, and an anodic peak at 0.50 V is related to the transition from the LixSi phase to amorphous Si according to Eq (1) and Eq (2).

Figure 5. (a) CV curves of the Si@MnO2@rGO-50°C at 0.1 mV s−1 between 0.1 V and 3.2 V. (b) Galvanostatic charge–discharge curves of the Si@MnO2@rGO-50°C at 0.1 A g−1. (c) Long-term cycling performance of the Si, Si@MnO2-50°C, Si@rGO-50°C, Si@MnO2@rGO-20°C, Si@MnO2@rGO-50°C, and Si@MnO2@rGO-160°C at 0.1 A g−1. (d) Rate performance of the Si@MnO2@rGO-50°C. (e) Long-term cycling performance of the Si@MnO2@rGO-50°C at 1 A g−1.
The redox peaks of Si coincided with those reported previously [10][11][12]. In addition, the reversible redox peaks at 1.23 and 0.36 V, respectively, were consistent with the lithiation and delithiation reactions of MnO2, according to Equation (3). When the cathodic peak was 0.36 V, Li was inserted into the anode to form LiO2 and MnO2 was reduced to Mn. An anodic peak at 1.23 V was related to the charging process of the lithium ion battery; Mn can facilitate the decomposition of LiO2.
In the subsequent scans, the redox peaks were largely coincident, which was attributed to the stability of the Si@MnO2@rGO-50°C, indicating the good electrochemical reversibility for lithium-ion batteries.
Figure 5b shows the galvanostatic discharge/charge curves of Si@MnO2@rGO-50°C at 0.1 A g−1 over the range between 0.1 V and 3.2 V. In the first cycle, the initial discharge capacity was 1378.15 mAh g−1. In the second, third, fifth, and 10th cycles, the discharge specific capacities were 1279.21, 1208.02, 1150.74, and 1093.70 mAh g−1, respectively. In the subsequent cycles, the charge and discharge curves basically overlapped, which indicated good capacity retention. In addition, the plateau of Si at 0.5 V in the figure was consistent with the CV results.
Figure 5c shows the cycling performances of the anode (Si, Si@MnO2-50°C, Si@MnO2@rGO-20°C, Si@MnO2@rGO-50°C, and Si@MnO2@rGO-160°C) at a current density of 0.1 A g−1. Of these samples, Si@MnO2@rGO-50°C showed the best cycling performance. The initial discharge specific capacity of Si was 1855.62 mAh g−1.  The initial specific capacity of Si@rGO-50°C was 1180.41 mAh g−1; after 150 cycles, the capacity was maintained at 543.84 mAh g−1 (Figure 5c).  Interestingly, the structure of rGO encapsulating the sea urchin-like Si@MnO2-50°C mitigated the volume-change effects of Si nanospheres, improving the electrical conductivity and contributing to the high capacity when comparing the Si and Si@MnO2-50°C. A comparison of the rate and cycling performances of the sea urchin-like Si@MnO2@rGO-50°C, at different various current densities, is shown in Figure 5d. At current densities of 0.1, 0.2, 0.5, 1, and 2 A g−1, the specific charging capacities were 1323.87, 971.85, 701.12, 491.85, and 272.47 mAh g−1, respectively. Furthermore, when the current density rose again to 0.1 A g−1, the specific charging capacity recovered to 981.68 mAh g1 and the capacity retention rate was 74.2%. Compared to previous reports of Si and Si@rGO, Si@MnO2@rGO-50°C demonstrated a long lifetime and high capacity retention rates [13][14]. Furthermore, to evaluate the cycling stability at high current densities, Si@MnO2@rGO-50°C was tested at 1 Ah g−1 (Figure 5e). In the first charge/discharge cycle, the specific capacity was 1446.85 mAh g−1; after 1000 cycles, the specific capacity was 1282.72 mAh g−1 with a coulombic efficiency of 99.4%. The composite material also showed excellent cycling performance at high currents. The decrease in capacity fluctuations at the first 80 cycles was due to the irreversible formation of SEI film on the materials’ surface. 
EIS measurements were carried out to further investigate the electrochemical mechanism shown in Figure 6Figure 6a shows the EIS plots and corresponding fitting plots of Si, Si@MnO2-50°C, and Si@MnO2@rGO-50°C, respectively. Additionally, the inset shows the data equivalent circuit diagrams. The EIS diagram consists of a semicircle in the mid-high-frequency region and a diagonal line in the low-frequency region. Here, RCT in the mid-frequency region corresponds to the charge-transfer impedance and WO in the low-frequency region corresponds to the diffusion of Li+ inside the electrode material [15][16]. Figure 6b shows the EIS plots and corresponding fitting plots of Si@MnO2@rGO-50°C before and after 150 cycles at 0.1 A g−1; cycled Si@MnO2@rGO-50°C was fitted with the same equivalent circuit model in Figure 6a. The RCT (72.7 Ω) was lower after cycling, which implies the charge transfer impedance decreased substantially, demonstrating the stable SEI film and good structure stability during cycling of Si@MnO2@rGO-50°C.
Figure 6. (a) The electrochemical impedance spectra of Si, Si@MnO2-50°C, and Si@MnO2@rGO-50°C. The inset shows the equivalent circuit model. (b) The electrochemical impedance spectra of the Si@MnO2@rGO-50°C before and after cycling.
Figure 7a,b shows the cross-sectional SEM images of Si before and after 150 cycles at 0.1 A g−1. After cycles, the volume expansion was ~323% (from 5.3 μm to 22.41 μm). The huge change in volume expansion of Si may have contributed to the intercalation and deintercalation of lithium ions during cycles and the exfoliation of the active materials. The mechanism of the Si electrode reaction was further illustrated in Figure 7e. The lithiation of Si resulted in the formation of an amorphous silicon–lithium alloy (LixSi) (0.16 V). Then, the amorphous LixSi was transformed to crystalline Li15Si4 (0.1 V). The delithiation process involves the transformation from crystalline Li15Si4 to amorphous LixSi (0.24 V) and, finally, to amorphous Si (0.5 V) [ 58 , 59 , 60 ]。

Figure 7. (a)&(b) Cross-sectional SEM images of Si before and after 150 cycles. (c)&(d) Cross-sectional SEM images of Si@MnO2@rGO-50℃ before and after 150 cycles. (e) The mechanism of intercalation and deintercalation of lithium ions.

References

  1. Chen, S.; Chen, Z.; Xu, X.Y.; Cao, C.B.; Xia, M.; Luo, Y.J. Scalable 2D Mesoporous Silicon Nanosheets for High-Performance Lithium-Ion Battery Anode. Small 2018, 14, 1703361.
  2. Zhao, T.M.; Han, Y.C.; Qin, L.N.; Guan, H.Y.; Xing, L.L.; Li, X.J.; Xue, X.Y.; Li, G.L.; Zhan, Y. Bidirectional modulation of neural plasticity by self-powered neural stimulation. Nano Energy 2021, 85, 106006.
  3. Zhao, Y.; He, J.F.; Dai, M.Z.; Zhao, D.P.; Wu, X.; Liu, B.D. Emerging 2 electrode materials assembled using nanosheets for flexible and foldable energy storage devices. J. Energy Chem. 2020, 45, 67–73.
  4. Liu, H.D.; Hu, Z.L.; Su, Y.Y.; Ruan, H.B.; Hu, R.; Zhang, L. MnO2 nanorods/3D-rGO composite as high performance anode materials for Li-ion batterie. Appl. Surf. Sci. 2017, 392, 777–784.
  5. Li, Y.; Ye, D.X.; Liu, W.; Shi, B.; Guo, R.; Pei, H.J.; Xie, J.Y. A three-dimensional core-shell nanostructured composite of polypyrrole wrapped MnO2/reduced graphene oxide/carbon nanotube for high performance lithium ion batteries. J. Colloid Interf. Sci. 2017, 493, 241–248.
  6. Xu, P.; Chen, H.Y.; Zhou, X.; Xiang, H.F. Gel polymer electrolyte based on PVDF-HFP matrix composited with rGO-PEG-NH2 for high-performance lithium ion battery. J. Membr. Sci. 2021, 617, 118660.
  7. Li, J.B.; Ding, Z.B.; Li, J.L.; Wang, C.Y.; Pan, L.K.; Wang, G.X. Synergistic coupling of NiS1.03 nanoparticle with S-doped reduced graphene oxide for enhanced lithium and sodium storage. Chem. Eng. J. 2020, 407, 127199.
  8. Fang, Y.Z.; Hu, R.; Liu, B.Y.; Zhang, Y.Y.; Zhu, K.; Yan, J.; Ye, K.; Cheng, K.; Wang, G.L.; Cao, D.X. MXene-derived TiO2/reduced graphene oxide composite with an enhanced capacitive capacity for Li-ion and K-ion batteries. J. Mater. Chem. A 2019, 7, 5363–5372.
  9. Yao, W.Q.; Wu, S.B.; Zhan, L.; Wang, Y.L. Two-dimensional Porous Carbon-coated Sandwich-like Mesoporous SnO2/Graphene/Mesoporous SnO2 Nanosheets towards High-Rate and Long Cycle Life Lithium-ion Batteries. Chem. Eng. J. 2018, 361, 329–341.
  10. Chae, S.; Choi, S.H.; Kim, N.; Sung, J.; Cho, J. Integration of Graphite and Silicon Anodes for the Commercialization of High-Energy Lithium-Ion Batteries. Angew. Chem. Int. Ed. 2019, 59, 110–135.
  11. Li, Z.H.; Zhang, Y.P.; Liu, T.F.; Gao, X.H.; Li, S.Y.; Ling, M.; Liang, C.D.; Zheng, J.C.; Lin, Z. Silicon Anode with High Initial Coulombic Efficiency by Modulated Trifunctional Binder for High-Areal-Capacity Lithium-Ion Batteries. Adv. Energy Mater. 2020, 10, 1903110.
  12. Keller, C.; Desrues, A.; Karuppiah, S.; Martin, E.; Alper, J.; Boismain, F.; Villevieille, C.; Herlin-Boime, N.; Haon, C.; Chenevier, P. Effect of Size and Shape on Electrochemical Performance of Nano-Silicon-Based Lithium Battery. Nanomaterials 2021, 11, 307.
  13. Zhao, Z.Q.; Cai, X.; Yu, X.Y.; Wang, H.Q.; Li, Q.Y.; Fang, Y.P. Zinc-assisted Mechanochemical Coating of Reduced Graphene Oxide Thin Layer on Silicon Microparticles for Efficient Lithium-ion Battery Anodes. Sustain. Energy Fuels. 2019, 3, 1258–1268.
  14. Li, X.; Bai, Y.S.; Wang, M.S.; Wang, G.L.; Ma, Y.; Huang, Y.; Zheng, J.M. Dual Carbonaceous Materials Synergetic Protection Silicon as a High-Performance Free-Standing Anode for Lithium-Ion Battery. Nanomaterials 2019, 9, 650.
  15. Chen, Y.X.; Yan, Y.J.; Liu, X.L.; Zhao, Y.; Wu, X.Y.; Zhou, J.; Wang, Z.F. Porous Si/Fe2O3 Dual Network Anode for Lithium–Ion Battery Application. Nanomaterials 2020, 10, 2331.
  16. Zhang, Y.; Hu, K.; Zhou, Y.L.; Xia, Y.B.; Yu, N.F.; Wu, G.L.; Zhu, Y.S.; Wu, Y.P.; Huang, H.B. A Facile, One-Step Synthesis of Silicon/Silicon Carbide/Carbon Nanotube Nanocomposite as a Cycling-Stable Anode for Lithium Ion Batteries. Nanomaterials 2019, 9, 1624.
  17. Han, N.; Li, J.J.; Wang, X.C.; Zhang, C.L.; Liu, G.; Li, X.H.; Qu, J.; Peng, Z.; Zhu, X.Y.; Zhang, L. Flexible Carbon Nanotubes Confined Yolk-Shelled Silicon-Based Anode with Superior Conductivity for Lithium Storage. Nanomaterials 2021, 11, 699.
  18. Kwon, H.J.; Hwang, J.Y.; Shin, H.J.; Jeong, M.G.; Chung, K.Y.; Sun, Y.K.; Jung, H.G. Nano/Microstructured Silicon–Carbon Hybrid Composite Particles Fabricated with Corn Starch Biowaste as Anode Materials for Li-Ion Batteries. Nano Lett. 2019, 20, 625–635.
  19. Argyropoulos, D.P.; Zardalidis, G.; Giotakos, P.; Daletou, M.; Farmakis, F. Study of the Role of Void and Residual Silicon Dioxide on the Electrochemical Performance of Silicon Nanoparticles Encapsulated by Graphene. Nano Lett. 2021, 11, 2864.
More
Information
Contributor MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register :
View Times: 578
Revisions: 3 times (View History)
Update Date: 24 Feb 2022
1000/1000
Video Production Service