Fabrication of carbon-coated V2O5-x nanoparticles by plasma-enhanced chemical vapor deposition for high-performance aqueous zinc-ion battery composite cathodes

Canglong Li Tao Liao Dongping Chen Tiancheng You Xiaozhi Jiang Minghan Xu Huaming Yu Gang Zhou Guanghui Li Yuejiao Chen

Citation:  Canglong Li, Tao Liao, Dongping Chen, Tiancheng You, Xiaozhi Jiang, Minghan Xu, Huaming Yu, Gang Zhou, Guanghui Li, Yuejiao Chen. Fabrication of carbon-coated V2O5-x nanoparticles by plasma-enhanced chemical vapor deposition for high-performance aqueous zinc-ion battery composite cathodes[J]. Chinese Chemical Letters, 2025, 36(12): 110557. doi: 10.1016/j.cclet.2024.110557 shu

Fabrication of carbon-coated V2O5-x nanoparticles by plasma-enhanced chemical vapor deposition for high-performance aqueous zinc-ion battery composite cathodes

English

  • Virtually all portable electronic devices, from cell phones and laptops to hybrid vehicles, rely on rechargeable battery technology [1,2]. Nevertheless, the widely used lithium-ion batteries (LIBs) have significant safety and environmental issues due to the formation of lithium dendrites, the high reactivity of lithium metal, and the toxicity of organic electrolytes [36]. Against this background, aqueous zinc-ion batteries (ZIBs) are suitable for large-scale energy storage, owing to their unique advantages, such as high theoretical capacity (820 mAh/g and 5855 mAh/cm3), low redox potential (−0.76 V vs. the standard hydrogen electrode), relatively high stability in water, and abundant resources [7,8]. Unfortunately, ZIBs still lack cathode materials which are structurally stable and have high capacity [9]. Currently, host materials for Zn2+ ions are predominantly composed of oxygen compounds that feature either a layered or tunnel-like structure. Notable examples include manganese oxide and vanadium oxide, as well as the versatile Prussian blue and its various derivatives [10]. Vanadium-based materials, compared to other cathode material systems, hold great potential as cathode materials for AZIBs, owing to their high theoretical zinc storage capacity, multi-valence of vanadium, and low raw material costs [11,12]. However, they face challenges such as inherently low electronic conductivity, incompatibility between interlayer spacing and Zn2+/Zn(H2O)n2+, and strong electrostatic interactions with zinc ions [13]. These issues lead to slow Zn2+ diffusion rates, dissolution of cathode materials in the electrolyte, and collapse of the host structure, severely hindering their utilization in large-scale applications.

    To address the limitation of rate performance and cycling stability, numerous strategies have been proposed, such as the application of nanostructured materials, amorphous phases, and the fabrication of protective coatings [14]. In terms of enhancing electronic conductivity, blending cathode materials with highly conductive carbon materials is the most cost-effective measure [15]. However, these common conductive materials like carbon black and graphene are low-dimensional. Due to the limited conductive contact with active materials, it is difficult to form a continuous and stable conductive network [16]. Suppressing the dissolution/collapse of cathode materials while improving electronic and ionic transport properties poses a significant challenge. In recent years, cathode materials with carbon coatings synthesized through methods such as mechanical ball milling, wet chemical methods, and thermal plasma have received widespread attention in various battery systems. The composite cathode materials prepared exhibit a stable structure which can inhibit adverse reactions of active particles, significantly improve the reaction kinetics, and achieve structural stability along with controllable volume changes [17]. However, complex multi-step processing or the use of toxic raw materials makes these methods environment unfriendly and less cost-effective [18]. Therefore, there is an urgent need for a simple and economical approach to prepare composite vanadium-based cathode materials with high ionic/electronic conductivity and structural stability.

    Herein, we successfully constructed novel carbon-coated V2O5-x nanoparticles (V2O5-x@C) via a simple plasma-enhanced chemical vapor deposition (PECVD) method, which significantly boosted the superior cycling stability and reaction kinetics as the cathode for AZIBs. As demonstrated through various electrochemical and material characterizations, V2O5-x@C cathode combines the superior structural characteristics of V2O5 nanoparticles and uniform conductive carbon coating layer, more oxygen vacancies, larger specific surface area, and higher zinc storage capacity, which guarantees fast electron and ion transfer kinetics and structural stability of the electrode. Impressively, a large specific surface area (4.1615 m2/g) and excellent stability of electrode in the aqueous electrolyte can be well achieved, which can be ascribed to the unique composite structure. Benefiting from rapid reaction kinetics and cycling stability, V2O5-x@C cathode delivers a high initial specific capacity of 149.2 mAh/g, with an excellent capacity retention of 83.8% at 5 A/g after 1000 cycles. Even at a high current density of 10 A/g, the specific discharge capacity of V2O5-x@C cathode still can remain at 80.2 mAh/g after 1000 cycles, with an attenuation of approximately 25.8%. This research suggests that directly constructing carbon-coating layer on cathode materials through PECVD method could vastly motivate the development of the design principle for high-performance cathode materials towards superior electrochemical energy storage systems.

    The synthetic procedure of V2O5-x@C cathode is illustrated in Fig. 1a. The V2O5 nanoparticles were coated with a uniform amorphous carbon layer via PECVD using methane (CH4) as the carbon precursor. Meanwhile, the vanadium of the +5 oxidation state was partially downgraded to +4 and +3, concurrently with the emergence of oxygen vacancies which ensured electroneutrality, culminating in the fabrication of the V2O5-x@C cathode. This is because the ion bombardment generated by gas ionization in the plasma promotes the reduction of vanadium, thereby reducing their valence. It is worth noting that unlike traditional wet chemical coating procedures, PECVD method can achieve uniform coatings with ideal carbon layer thickness without the need for additional work steps or the generation of by-products. The morphology and microstructure of the V2O5 and V2O5-x@C were characterized by scanning electron microscope (SEM) and transmission electron microscopy (TEM) characterizations. As shown in Fig. S1 (Supporting information), the pristine V2O5 exhibits irregular particle distribution with sizes accompanied by serious agglomeration. Under the action of high-energy plasma, the irregular shape of V2O5 nanoparticles undergoes a transformation, and the V2O5-x@C reveals the uniform spherical nanoparticle morphology (Fig. 1b). Noticeably, by strictly controlling the degree of reduction, V2O5-x@C can still maintain good dispersibility without significant agglomeration, which can increase the electrochemical active surface area of the cathode material and be beneficial for improving the capability to store zinc.

    Figure 1

    Figure 1.  (a) Schematic illustration of the synthesis process of the V2O5-x@C cathode material. (b) SEM images of V2O5-x@C and (c) EDS mapping of C, V, and O elements. (d–f) TEM images of the V2O5-x@C.

    In addition, the energy dispersive spectroscopy (EDS) elemental mapping images in Fig. 1c confirm the uniform dispersion of the O and V elements among the entire spherical particles, and C element is distributed on the outer surface of nanoparticles, while nearly no C element is found in the pure V2O5 (Figs. S2 and S3 in Supporting information), further evidencing the homogeneous coverage of carbon coating on the surface of V2O5-x@C. A closer observation of transmission electron microscopy (TEM) clearly shows a thin and uniform coating layer on the surface of V2O5-x@C nanoparticles and the average thickness of the carbon coating layer is around 20 nm (Fig. 1d). Combined with the high-resolution TEM images in Figs. 1e and f, the defined lattice fringes can be observed, in which the interplanar spacing about 0.57 nm and 0.34 nm can be indexed to the (001) plane of V2O5 and (101) plane of VO2, respectively [19].

    Consistent results can also be observed in X-ray diffraction (XRD) pattern in Fig. 2a, the pattern of V2O5-x@C exhibits the characteristic diffraction peaks of VO2 (JCPDS No. 76-0677), except the characteristic diffraction peaks of V2O5 (JCPDS No. 09-0387). Moreover, unlike the brownish yellow V2O5 powder, V2O5-x@C powder displays a uniform brownish black color, as shown in the insets of Fig. 2a. Similar result can be further proved by the Raman measurement (Fig. 2b). Raman spectra reveal that V2O5 and V2O5-x@C display similar stretching and bending vibrations of chemical bonds in the region of 100–1200 cm−1, indicating that the V2O5-x@C composite cathode still retains the V=O and V-O-V chemical bonds of V2O5 (Fig. 2c). Remarkably, the peak at 877 cm−1 represents the C-O-C band of carbon, which correspond to the carbon layer on the surface of V2O5-x@C [20,21]. These results further confirm that during the PECVD process, a portion of V5+ in V2O5 powder is reduced to low valence states and exists on the surface of the nanoparticles.

    Figure 2

    Figure 2.  (a) XRD patterns (Insets show photographic images of V2O5-x@C and V2O5 powders), (b) Raman shift, (c) normalized Raman spectroscopy with an energy range of 650–750 cm−1, (d) EPR spectra, (e) TGA curves, and (f) BET adsorption isotherm, of the synthesized V2O5-x@C and pure V2O5 powders.

    We deemed that a large number of oxygen vacancies will be generated during the transition of V element from +5 valence to low valence states, so Raman spectroscopy and electron paramagnetic resonance (EPR) spectra were conducted to prove this conjecture. Combining the results of Figs. 2b and c, it can be found that the Raman spectrum of V2O5-x@C has a slight blue shift relative to the pristine V2O5, which can be considered as the effect of existent oxygen vacancies [2224]. In order to determine the content of oxygen vacancies, we obtained the EPR spectra of pristine V2O5 and V2O5-x@C samples (Fig. 2d). A clear oxygen vacancy signal at a typical g-factor equal to 2.006 is observed in the EPR spectra, demonstrating the content of oxygen vacancies can be significantly enhanced through PECVD treatment. Plentiful oxygen vacancies can stimulate cathode adsorption affinity which would contribute to the improvement of storage capacity and electrochemical performance. Furthermore, the carbon loading in the V2O5-x@C was characterized by thermogravimetric analysis (TGA) [25]. The 0.66% weight loss of V2O5-x@C between 0 and 350 ℃ could be ascribed to the removal of carbon coating (Fig. 2e). It is noteworthy that although the carbon content is low, the PECVD treatment process not only covers the surface of V2O5 with a carbon layer, but also significantly improves the uniformity of nanoparticles and suppresses agglomeration, thereby significantly increasing the surface area (4.1615 m2/g for V2O5-x@C, and 2.8695 m2/g for V2O5) (Fig. 2f), which is conducive to the diffusion and transport of zinc ions [2628].

    To determine the valence variation of V states and the generation of carbon layer after reduction during PECVD process, the V2O5 and V2O5-x@C electrodes were evaluated via X-ray photoelectron spectrometer (XPS) measurement. As depicted in Fig. 3a, V 2p and O 1s peaks are observed in the XPS data of both V2O5 and V2O5-x@C electrodes, while an obvious C 1s peak can be found in the data of V2O5-x@C electrode. Besides, in the C 1s spectrum, the three peaks are concentrated at 284.8, 286.7 and 288.8 eV corresponding to the C-C/C-H, C-O and C=O bonds, respectively, proving that uniform and stable carbon coating is generated on the surface of the cathode materials during the preparation process (Fig. 3b) [2931]. Only the peaks of V5+ present at 517.6 and 525.0 eV in the V 2p spectrum of pristine V2O5, which is in good consistent with the V2O5 crystal structure (Fig. 3c). Conversely, the V 2p spectrum of V2O5-x@C electrode displays three spin-orbit peaks, which can be attributed to V5+ (517.6 and 525.0 eV), V4+ (516.5 and 523.4 eV), and a low content of V3+, indicating that PECVD process can lead to the formation of a mixed valence state in V2O5 (Fig. 3d) [19,32]. This facilitates an increase in electrical conductivity through electron hopping among V3+, V4+, and V5+. Concurrently, the formation of oxygen vacancies helps to preserve electroneutrality, which in turn enhances ionic conductivity and improves charge-transfer kinetics [33,34]. Moreover, benefiting from the carbon layer coated on vanadium-based oxide nanoparticles, the structure and interface stability of the electrode in aqueous electrolyte have been significantly improved [35,36]. After soaking V2O5 and V2O5-x@C electrodes in the electrolytes for 4 days, it can be observed that the color of the electrolyte within soaked V2O5 electrode has significantly turned yellow, which is due to the continuous dissolution of V2O5 in the electrolyte (Fig. 3e). The optical image of the V2O5 electrode surface also shows a large number of pores and uneven surfaces caused by the dissolution of the cathode material (Fig. 3f). Contrarily, the V2O5-x@C electrode immersed in the electrolyte did not show significant dissolution, and there is still a dense distribution of active material on the surface of the electrode.

    Figure 3

    Figure 3.  (a) Wide-scan XPS spectra of V2O5-x@C and V2O5. (b) C 1s XPS spectra of V2O5-x@C. V 2p XPS spectra of (c) V2O5 and (d) V2O5-x@C. (e) Optical photographs of V2O5-x@C and V2O5 cathodes soaked in the electrolyte. (f) Optical photographs of the surfaces of V2O5-x@C and V2O5 cathodes after soaking in the electrolyte for 4 days.

    To investigate the electrochemical capacity retention and cycling stability, the performance and coulombic efficiency of V2O5 and V2O5-x@C electrodes are tested during extended cycling (Fig. S4 in Supporting information). Firstly, two pairs of redox peaks are found in the CV curves which are shown in Fig. 4A, indicating the Zn2+ intercalation/deintercalation reactions. Meanwhile, the reduction of the potential difference between the redox peaks of V2O5-x@C electrode demonstrates enhanced electrochemical activity and reversibility. In addition, the promotion of V2O5-x@C on electron transfer is characterized by Nyquist impedance spectra (Fig. 4B). The Zn//V2O5-x@C full cell delivers a far smaller interface impedance than that of Zn//V2O5 cell, indicating the facilitated charge transfer process and good interfacial compatibility at the anode/electrolyte interface [37,38]. The rate performance is given in Figs. 4CE. The enhanced electrochemical kinetics demonstrates that the V2O5-x@C electrode is highly effective in facilitating the Zn2+ diffusion process.

    Figure 4

    Figure 4.  Electrochemical performance of V2O5-x@C and V2O5 electrodes. (A) CV curves at a scan rate of 1 mV/s. (B) Nyquist plots. (C) Rate capacity at 0.5–10 A/g. (D, E) Discharge curves at different current densities. Cycle performance and coulombic efficiency at (F) 5 A/g and (G) 10 A/g. (H) Galvanostatic charge/discharge profiles at 10 A/g.

    Zn//V2O5 cell exhibits lower initial specific capacity, and with increasing current density, the specific capacity significantly decreases, mainly due to lower electronic conductivity and limited diffusion of ions in active materials. On the contrary, Zn//V2O5-x@C full cell delivers higher specific capacities at various current densities, and remains 277.6 mAh/g when returned to 1 A/g. It is concluded that V2O5-x@C composite cathodes exhibit a highly stable structure and faster reaction kinetics during the current fluctuation, which can lead to an extremely stable performance during long-term cycling (Figs. 4FH) [39]. V2O5-x@C composite cathode shows a retained capacity of 83.8% upon 1000 cycles at 5 A/g, even at a higher current density of 10 A/g, the specific discharge capacity of V2O5-x@C cathode still can remain at 80.2 mAh/g after 1000 cycles. The cyclic stability of V2O5-x@C composite cathode is also much superior to most reported works under various testing conditions (Table S1 in Supporting information). Noticeably, according to the interfacial impedance after cycling (Fig. S5 in Supporting information), the interface stability is better in V2O5-x@C composite cathode [40]. Moreover, it is generally approved that the Raman signal at around 980 cm−1 from SO42− can reflect the concentration change of zinc ions to some extent [41]. As shown in Fig. S6 (Supporting information), a constant Raman signal is observed at the interface of V2O5-x@C cathode during cycling, testifying optimized ion flux and fast transfer dynamics of Zn ions at the interface. Importantly, the Zn//V2O5-x@C full cell shows better self-discharge performance, which is evidenced by higher capacity retention over 34 h (Fig. S7 in Supporting information). The electrochemical performance is greatly optimized owing to the improved electrical conductivity, enhanced structural stability, and the increased specific surface area of the V2O5-x@C composite cathode.

    It is widely recognized that lattice and phase transformations are crucial during the charge/discharge process [4244]. To elucidate these structural dynamics, ex situ XRD and XPS analyses were conducted on V2O5-x@C cathode, shedding light on its evolution during the intercalation/deintercalation of Zn2+ ions. Specifically, the XRD diffraction patterns at varying voltages provide a visual representation of the structural transition (Figs. 5ac). During the discharge phase, as the voltage decreased from 1.6 V to 0.4 V, a slight shift of the diffraction peak can be observed moving towards lower angles. This deviation is indicative of Zn2+ ions inserting into the electrode, leading to an expansion of the interlayer spacing [45]. Conversely, during the charging phase, the peak reverts to its initial angle, corresponding to the desertion process of Zn2+ ions. Notably, throughout this charge/discharge process, no phase transition or byproduct formation were detected. This observation confirms that the phase composition of V2O5-x@C remains stable and unaltered, underscoring its robustness and reliability as an electrode material in the application of AZIBs [46].

    Figure 5

    Figure 5.  (a–c) Ex situ XRD pattern of V2O5-x@C electrodes during the charge/discharge process in the first cycle. (d–f) High-resolution V 2p spectra of V2O5-x@C electrodes under different voltages. (g) Diagram of the alteration in the lattice structure during the zinc ions intercalation/deintercalation process.

    Additionally, the variation trend of the V 2p peak during the charge and discharge processes is depicted in Figs. 5df. As the results mentioned in Fig. 3d shows, V5+ is dominant in the pristine V2O5-x@C electrode, with a minor presence of V3+ and V4+. Upon discharging to 0.4 V, there is a significant reduction in the proportion of V5+, coinciding with an increase in the percentages of V3+ and V4+. This shift implies the insertion of Zn2+ ions, which leads to a reduction in the valency of vanadium. Instead, when the cell was charged to 1.6 V, the valence state of vanadium reverts to its original state, signifying that the Zn2+ ions can be efficiently deserted [47]. The alterations in the lattice structure throughout the Zn2+ intercalation and deintercalation processes are clearly presented in Fig. 5g. The V2O5-x@C nanoparticles are tightly combined with a carbon coating layer, which remains stable during both the Zn2+ insertion and extraction processes. A marked reduction in the content of Zn2+ occurs after the charge process. According to the above analysis results, V2O5-x@C electrode exhibits the high reversibility of the lattice structure, with no indication of phase transition or the formation of any byproducts, which is in favor of remarkably improving the electrochemical performance.

    In conclusion, a novel V2O5-x@C composite cathode composed of oxygen-vacancy-abundant V2O5-x nanoparticles and carbon layer with high conductivity was successfully constructed via a simple PECVD method. Uniformly dispersed spherical V2O5-x nanoparticles significantly increase the surface area, providing more active sites for electrochemical reactions. The rich oxygen vacancies and the conductive carbon layer coating ensure rapid electron transfer and greatly enhance the diffusion kinetics of zinc ions. Additionally, the composite structure of the cathode guarantees the structural stability of the electrode during the repeating charge/discharge process and the interface reliability in the aqueous electrolyte. Benefiting from the fast electron/ion kinetics, the novel V2O5-x@C composite cathode achieves ideal surface area (4.1615 m2/g) and excellent cycling stability, as demonstrated via various material characterizations and electrochemical analyses. Impressively, V2O5-x@C electrode exhibits the comprehensive performance of excellent rate and cycle stability, along with minimal structural alteration upon Zn2+ intercalation and high reversibility. The Zn//V2O5-x@C cell achieves a high initial specific capacity of 149.2 mAh/g at 5 A/g, possessing a capacity retention of 83.8% over 1000 cycles. Such stabilized crystal structure and enhanced ion diffusion coating of V2O5-x@C cathode herald promising prospects for the development of high-performance AZIBs.

    The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

    Canglong Li: Software, Methodology. Tao Liao: Writing – review & editing. Dongping Chen: Validation. Tiancheng You: Investigation. Xiaozhi Jiang: Investigation. Minghan Xu: Investigation, Formal analysis. Huaming Yu: Writing – original draft, Visualization, Supervision. Gang Zhou: Supervision, Software. Guanghui Li: Project administration. Yuejiao Chen: Writing – review & editing, Supervision, Funding acquisition.

    This research was financially supported by the National Natural Science Foundation of China (No. 52377222) and Natural Science Foundation of Hunan Province (No. 2023JJ20064).

    Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.cclet.2024.110557.


    1. [1]

      D. Sun, Z. Sun, D. Yang, et al., EcoEnergy 1 (2023) 375–404. doi: 10.1002/ece2.22

    2. [2]

      Q.C. Wang, S.H. Tang, Z.Q. Wang, et al., Adv. Funct. Mater. 33 (2023) 2307390. doi: 10.1002/adfm.202307390

    3. [3]

      H.M. Yu, C.D. Lv, C.S. Yan, et al., Small Methods 8 (2023) 2300758.

    4. [4]

      D.P. Chen, P. Qing, F.C. Tang, et al., Mater. Today Energy 33 (2023) 101272. doi: 10.1016/j.mtener.2023.101272

    5. [5]

      F. Chen, Z.L. Xu, Microstructures 2 (2022) 2022012.

    6. [6]

      Z. Wang, J. Xue, Z. Song, et al., Chin. Chem. Lett. 35 (2024) 109489. doi: 10.1016/j.cclet.2024.109489

    7. [7]

      H. Yu, Z. He, D. Chen, et al., Energy Rev. 4 (2025) 100107. doi: 10.1016/j.enrev.2024.100107

    8. [8]

      J. Zhou, H. Yu, P. Qing, et al., J. Colloid Interface Sci. 678 (2024) 772–782.

    9. [9]

      Z. Wang, L. Yang, C. Xu, et al., Green Carbon 1 (2023) 193–209. doi: 10.1016/j.greenca.2023.12.001

    10. [10]

      D.S. Liu, Z.Y. Zhang, Y.F. Zhang, et al., Angew. Chem. Int. Ed. 62 (2023) e202215385. doi: 10.1002/anie.202215385

    11. [11]

      H.L. Cao, Z.Y. Zheng, P. Norby, X.X. Xiao, S. Mossin, Small 17 (2021) 2100558. doi: 10.1002/smll.202100558

    12. [12]

      M. Chen, S.C. Zhang, Z.G. Zou, et al., Rare Met. 42 (2023) 2868–2905. doi: 10.1007/s12598-023-02303-2

    13. [13]

      S.W. Wang, H. Zhang, K. Zhao, et al., Carbon Energy 5 (2023) e330. doi: 10.1002/cey2.330

    14. [14]

      S. Huang, K. Li, Z. He, et al., J. Alloy. Compd. 1005 (2024) 176193. doi: 10.1016/j.jallcom.2024.176193

    15. [15]

      A. Habibi, M.R. Mousavi, M. Yasoubi, Z. Sanaee, S. Ghasemi, Mater. Sci. Semicond. Process. 149 (2022) 106901. doi: 10.1016/j.mssp.2022.106901

    16. [16]

      H.R. Zheng, D.N. Deng, X.R. Zheng, et al., Nano Lett. 24 (2024) 4672–4681. doi: 10.1021/acs.nanolett.4c01078

    17. [17]

      X.Y. Zhang, Y.R. Wang, B.I. Min, et al., Adv. Powder Technol. 32 (2021) 2828–2838. doi: 10.1016/j.apt.2021.06.003

    18. [18]

      F. Bu, Y. Gao, W.B. Zhao, et al., Angew. Chem. Int. Ed. 63 (2024) e202318496. doi: 10.1002/anie.202318496

    19. [19]

      F.Y. Chen, H.R. Luo, M. Li, et al., ACS Appl. Mater. Interfaces 14 (2022) 53677–53689. doi: 10.1021/acsami.2c14153

    20. [20]

      X. Zhang, H. Weng, Y. Miu, et al., Chem. Eng. J. 482 (2024) 148807. doi: 10.1016/j.cej.2024.148807

    21. [21]

      W. Liu, X. Liu, F. Ning, et al., J. Mater. Chem. A 12 (2024) 11883–11894. doi: 10.1039/d4ta00376d

    22. [22]

      H. Ge, W.P. Kong, R. Wang, et al., Opt. Lett. 48 (2023) 2186–2189. doi: 10.1364/ol.486417

    23. [23]

      P. Cui, Y.X. Zhang, Z. Cao, et al., Chem. Eng. J. 412 (2021) 128676. doi: 10.1016/j.cej.2021.128676

    24. [24]

      X. Liu, Y. Guo, F. Ning, et al., Nano Micro Lett. 16 (2024) 111. doi: 10.1007/s40820-024-01337-0

    25. [25]

      X.F. Liu, H. Chen, R. Wang, S.Q. Zang, T.C.W. Mak, Small 16 (2020) 2002932. doi: 10.1002/smll.202002932

    26. [26]

      P.J. De, J. Halder, S. Priya, A.K. Srivastava, A. Chandra, ACS Appl. Energy Mater. 6 (2023) 753–762. doi: 10.1021/acsaem.2c02979

    27. [27]

      S.Z. Deng, B.A. Xu, J.X. Zhao, C.W. Kan, X.L. Liu, Angew. Chem. Int. Ed. 63 (2024) e202401996. doi: 10.1002/anie.202401996

    28. [28]

      J. Chen, M. Chen, H. Chen, et al., Proc. Natl. Acad. Sci. U. S. A. 121 (2024) e2322944121. doi: 10.1073/pnas.2322944121

    29. [29]

      J. Ren, H.Y. Wu, W. Yan, et al., Ind. Chem. Mater. 2 (2024) 328–339. doi: 10.1039/d3im00111c

    30. [30]

      J. Cao, H. Wu, D. Zhang, et al., Angew. Chem. Int. Ed. 63 (2024) e202319661. doi: 10.1002/anie.202319661

    31. [31]

      S. Chen, J.L. Chen, W. Wang, et al., Energy Storage Mater. 68 (2024) 103348. doi: 10.1016/j.ensm.2024.103348

    32. [32]

      Q. Wang, H. Zhao, M. Chen, et al., Chem. Eng. J. 484 (2024) 149674. doi: 10.1016/j.cej.2024.149674

    33. [33]

      S. Zhan, Y. Guo, K. Wu, et al., Chem. Eur. J. 30 (2024) e202303211. doi: 10.1002/chem.202303211

    34. [34]

      M. Han, J. Yao, J. Huang, et al., Chin. Chem. Lett. 34 (2023) 107493. doi: 10.1016/j.cclet.2022.05.007

    35. [35]

      Z. He, H. Yu, M. Fu, et al., Energy Storage Mater. 70 (2024) 103469. doi: 10.1016/j.ensm.2024.103469

    36. [36]

      H.Z. Ren, J. Zhang, B. Wang, et al., Rare Met. 41 (2022) 1605–1615. doi: 10.1007/s12598-021-01892-0

    37. [37]

      Z.Q. He, H.M. Yu, D.P. Chen, et al., Chem. Eur. J. 30 (2024) e202400567. doi: 10.1002/chem.202400567

    38. [38]

      X.P. Ma, H.M. Yu, C.S. Yan, et al., J. Colloid Interface Sci. 664 (2024) 539–548. doi: 10.1016/j.jcis.2024.03.085

    39. [39]

      D.F. Sun, Z.J. Wang, T. Tian, et al., Rare Met. 43 (2024) 635–646. doi: 10.1007/s12598-023-02434-6

    40. [40]

      Y. Zhong, C. Cao, L. Zhao, M.O. Tadé, Z. Shao, Green Carbon 2 (2024) 94–100. doi: 10.62051/dgm5pf38

    41. [41]

      J.Q. Zhou, L.F. Zhang, M.J. Peng, et al., Adv. Mater. 34 (2022) 2200131. doi: 10.1002/adma.202200131

    42. [42]

      C.Y. Chang, S.L. Hu, T.T. Li, et al., Energy Environ. Sci. 17 (2024) 680–694. doi: 10.1039/d3ee03422d

    43. [43]

      L.S. Han, Y.M. Guo, F.H. Ning, et al., Adv. Mater. 36 (2024) 2308086. doi: 10.1002/adma.202308086

    44. [44]

      C. Wang, Y. Yang, S. Zhang, et al., Journal of Energy Storage 94 (2024) 112360. doi: 10.1016/j.est.2024.112360

    45. [45]

      R. Jin, Y. Fang, B. Gao, et al., Ind. Chem. Mater. (2024) 10.1039/D4IM00042K. doi: 10.1039/d4im00042k

    46. [46]

      Z. Bie, Z.Y. Jiao, X.X. Cai, et al., Adv. Energy Mater. 14 (2024) 2401002. doi: 10.1002/aenm.202401002

    47. [47]

      W. Ling, C.X. Nie, X.W. Wu, et al., ACS Nano 18 (2024) 5003–5016. doi: 10.1021/acsnano.3c11115

  • Figure 1  (a) Schematic illustration of the synthesis process of the V2O5-x@C cathode material. (b) SEM images of V2O5-x@C and (c) EDS mapping of C, V, and O elements. (d–f) TEM images of the V2O5-x@C.

    Figure 2  (a) XRD patterns (Insets show photographic images of V2O5-x@C and V2O5 powders), (b) Raman shift, (c) normalized Raman spectroscopy with an energy range of 650–750 cm−1, (d) EPR spectra, (e) TGA curves, and (f) BET adsorption isotherm, of the synthesized V2O5-x@C and pure V2O5 powders.

    Figure 3  (a) Wide-scan XPS spectra of V2O5-x@C and V2O5. (b) C 1s XPS spectra of V2O5-x@C. V 2p XPS spectra of (c) V2O5 and (d) V2O5-x@C. (e) Optical photographs of V2O5-x@C and V2O5 cathodes soaked in the electrolyte. (f) Optical photographs of the surfaces of V2O5-x@C and V2O5 cathodes after soaking in the electrolyte for 4 days.

    Figure 4  Electrochemical performance of V2O5-x@C and V2O5 electrodes. (A) CV curves at a scan rate of 1 mV/s. (B) Nyquist plots. (C) Rate capacity at 0.5–10 A/g. (D, E) Discharge curves at different current densities. Cycle performance and coulombic efficiency at (F) 5 A/g and (G) 10 A/g. (H) Galvanostatic charge/discharge profiles at 10 A/g.

    Figure 5  (a–c) Ex situ XRD pattern of V2O5-x@C electrodes during the charge/discharge process in the first cycle. (d–f) High-resolution V 2p spectra of V2O5-x@C electrodes under different voltages. (g) Diagram of the alteration in the lattice structure during the zinc ions intercalation/deintercalation process.

  • 加载中
计量
  • PDF下载量:  0
  • 文章访问数:  19
  • HTML全文浏览量:  2
文章相关
  • 发布日期:  2025-12-15
  • 收稿日期:  2024-08-26
  • 接受日期:  2024-10-16
  • 修回日期:  2024-09-16
  • 网络出版日期:  2024-10-18
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

/

返回文章