Next Article in Journal
Determination of Anthraquinone-Tagged Amines Using High-Performance Liquid Chromatography with Online UV Irradiation and Luminol Chemiluminescence Detection
Previous Article in Journal
Siloxene Nanosheets and Their Hybrid Gel Glasses for Broad-Band Optical Limiting
Previous Article in Special Issue
Efficient Hydrogen and Oxygen Evolution Catalysis Using 3D-Structured Nickel Phosphosulfide Nanosheets in Alkaline Media
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Bio-Template Synthesis of V2O3@Carbonized Dictyophora Composites for Advanced Aqueous Zinc-Ion Batteries

1
College of Materials and Advanced Manufacturing, Hunan University of Technology, Zhuzhou 412007, China
2
Jiangsu Co-Innovation Center of Efficient Processing and Utilization of Forest Resources, International Innovation Center for Forest Chemicals and Materials, College of Materials Science and Engineering, Nanjing Forestry University, Nanjing 210037, China
3
Hunan Key Laboratory of Applied Environmental Photocatalysis, Changsha University, Changsha 410022, China
4
Institute of Materials Science and Devices, School of Materials Science and Engineering, Suzhou University of Science and Technology, Suzhou 215009, China
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(5), 2147; https://doi.org/10.3390/molecules28052147
Submission received: 2 February 2023 / Revised: 21 February 2023 / Accepted: 22 February 2023 / Published: 24 February 2023
(This article belongs to the Special Issue Advanced Energy Storage Materials and Their Applications)

Abstract

:
In terms of new-generation energy-storing devices, aqueous zinc-ion batteries (AZIBs) are becoming the prime candidates because of their inexpensive nature, inherent safety, environmental benignity and abundant resources. Nevertheless, due to a restrained selection of cathodes, AZIBs often perform unsatisfactorily under long-life cycling and high-rate conditions. Consequently, we propose a facile evaporation-induced self-assembly technique for preparing V2O3@carbonized dictyophora (V2O3@CD) composites, utilizing economical and easily available biomass dictyophora as carbon sources and NH4VO3 as metal sources. When assembled in AZIBs, the V2O3@CD exhibits a high initial discharge capacity of 281.9 mAh g−1 at 50 mA g−1. The discharge capacity is still up to 151.9 mAh g−1 after 1000 cycles at 1 A g−1, showing excellent long-cycle durability. The extraordinary high electrochemical effectiveness of V2O3@CD could be mainly attributed to the formation of porous carbonized dictyophora frame. The formed porous carbon skeleton can ensure efficient electron transport and prevent V2O3 from losing electrical contact due to volume changes caused by Zn2+ intercalation/deintercalation. The strategy of metal-oxide-filled carbonized biomass material may provide insights into developing high-performance AZIBs and other potential energy storage devices, with a wide application range.

1. Introduction

Lithium-ion batteries (LIBs), with the advantages of mature preparation technology and high energy density, are viewed as the most alluring candidates for advanced portable and automotive electrical energy storage systems [1,2,3,4]. Nonetheless, the deficiency in lithium assets and unpredictable safety problems of flammable electrolytes have genuinely hindered the further development and wide application of LIBs [5,6]. Consequently, it is imperative to find alternatives to LIBs. Recently, aqueous metal-ion (Zn, Na, K, Mg, Ca, etc.) batteries have exhibited extraordinary prospects for applications in energy storage, owing to their fabulous quality of security [7]. Among them, rechargeable aqueous zinc-ion batteries (AZIBs) are considered to be one of the foremost promising energy storage devices on account of their plenteous sources, nontoxicity and high theoretical capacity (820 mAh g−1) [8,9,10,11].
Within the past few decades, various types of host materials, including MnO2, Prussian blue analogues, organic compounds and V-based materials, have been considered extensively as AZIB cathodes [8,12]. However, Zn/MnO2 aqueous batteries feature an underdeveloped rate capability and significant capacity fading because of the sluggish dynamics and phase transitions [13,14]. Although Prussian blue analogue cathodes presented superior rate capability and cyclability, their practical application is still enormously constrained by the problem of low specific capacity of below 100 mAh g−1 and producing O2 evolution under high working voltage [15]. For organic compound electrodes, the formed discharge products are easily dissolved in the aqueous electrolyte, leading to cycling performance degradation [16]. In contrast, Vanadium oxides, with an open-layered structure, are considered as potential candidates for AZIB cathodes, which could benefit from the multivalent transition effect of vanadium to furnish the extra losses and gains of electrons, thus resulting in high theoretical capacity (generally >300 mAh g−1) [17]. In addition, differences in the oxidation states and REDOX properties of vanadium lead to the formation of different types of layered vanadium oxides and their vanadate, which can be applied to temperature sensors, catalysis and various optical and electrochemical devices [18,19]. Currently, the commonly employed vanadium oxides for electrode materials include V2O3 [20], H2V3O8 [1], CaV4O9 [21], VO2 [22], V3O7∙H2O [23], Zn3V2O7(OH)2∙2H2O [24], V6O13 [2] and so on.
However, vanadium oxides generally suffer from limited electrochemical performances, including dramatic capacity decay and unsatisfactory rate capability due to surface element dissolution, inferior electric conductivity and slow reaction kinetics. Previous research has indicated that compositing with carbon-based materials [25,26] cannot only construct a conducting framework for electron transfer, thus ameliorating the rate performance, but also effectively mitigate the instability of the layered structure of vanadium oxides [5]. Pang et al. [1] reported that H2V3O8 nanowire/graphene materials prepared via the hydrothermal method exhibited better cycling stability and far higher current density compared to a pure H2V3O8 electrode. Hu et al. [27] demonstrated that uniformly encapsulating V2O3 nanoparticles in amorphous carbon nanosheets can effectively promote the penetration of the electrolyte, leading to rapid and durable potassium storage behavior in PIBs. Dai et al. [28] prepared V2O5@polyaniline cathode materials for AZIBs, exhibiting a high reversible capacity of 361 mAh g−1 at 0.1 A g−1 and a lifespan of 1000 cycles (93.8% capacity retention). Tamilselvan et al. [2] designed and fabricated a V6O13@carbon cloth with excellent electrical conductivity, exhibiting an initial specific capacity of 227 mAh g−1 and nearly 99% retention rate after 1000 cycles. Although the above-mentioned synthetic methods are well developed in small-batch production, they are still far from commercialization due to the complex synthetic approaches, high cost and unsustainable carbon sources [29,30,31]. Thus, seeking efficient and renewable raw materials for preparing functional carbon materials has gradually gained attention. As a carbon-rich precursor, biomass has been widely studied and applied in many fields due to its natural advantages of renewable and abundant resources, environmental friendliness and economy [32,33,34,35,36].
In this work, aiming to enhance the electrochemical performance of vanadium oxides as AZIB cathodes, V2O3@carbonized dictyophora (V2O3@CD) composites were synthesized via an evaporation-induced self-assembly technique by using NH4VO3 as metal sources and carbonized dictyophora as carbon sources. Carbonized dictyophora provides a carbon skeleton for V2O3 by retaining its unique rhombohedral structure, which not only effectively alleviates the volume expansion of the electrode, but also improves the intercalation/de-intercalation process of zinc ions. Therefore, an exceptionally high capacity of 289 mAh g−1 at a current density of 50 mA g−1, a long-life cycling stability with a capacity retention rate of 89% after 1000 cycles at 50 mA g−1 and an excellent rate capability (170 mAh g−1 at 1 A g−1) are achieved in the as-prepared V2O3@CD composites. The V2O3@CD cathode could offer tremendous possibilities for fast and durable zinc-ion storage.

2. Results

The crystalline structure of as-synthesized V2O3@CD composites was characterized by an XRD experiment. The corresponding XRD patterns (Figure 1a) manifest that the distinct narrow peaks of all products can be assigned to rhombohedral V2O3 (space group: R-3c, ICSD-01-071-0342) with the lattice parameters of a = 4.95 Å, b = 4.95 Å, c = 14.00 Å, α = 90°, β = 90° and γ = 120° [37,38]. It is worth mentioning that a low-intensity amorphous bump centered at approximately 24.5°, which can be attributed to the (002) plane of amorphous C, and the intensity of this amorphous C peak, debilitates with the diminishing of the C atom substance [27]. No other noticeable impurity phases were detected in the pattern, suggesting that V2O3 was successfully synthesized [39]. Due to the low content of V, the diffraction peaks of the VOCD-1 sample are the weakest among all the as-prepared composites. With a higher content of V, the VOCD-3 sample shows stronger diffraction peaks compared with the other two V2O3@CD composites, indicating better crystallinity, which is beneficial to the ion diffusion between the layers of the electrode.
The FT-IR spectra of V2O3@CD composites are illustrated in Figure 1b. The bending vibration of O-H and the stretching vibration of H-O-H can be observed at 1624 and 3405 cm−1, respectively, which is mainly because some water molecules are embedded between the layers and adsorbed on the surface of the composites [40]. The peaks at 805 and 533 cm−1 can be indexed to the symmetric and asymmetric stretching vibration of V-O-V [41]. Notably, the peak centered at 984 cm−1 is associated with the stretching vibration of the V3+=O, demonstrating the appearance of V2O3 [42]. The peaks at 2855 and 2932 cm−1 are related to the stretching vibration of C-H bonds [43], while the peak located at 1421 cm−1 is characteristic of bending vibration of a C-H bond [44]. The asymmetrical stretching vibration of C-O is around 2356 cm−1, aiming at CO2 adsorption on KBr, which is negligible [45]. The above FT-IR spectra demonstrate that the as-obtained composites are composed of V2O3 and amorphous carbon, which is in agreement with the XRD results.
XPS spectrum was employed to investigate the elemental valence and surface composition of V2O3/CD composites. As illustrated in Figure 1c–f, all the signal peaks can be assigned to C 1s, O 1s and V 2p in the survey spectra. The deconvoluted C 1s spectra of V2O3/CD composites revealed several components at 287.9, 286.3 and 284.8 eV, which corresponded to O-C=O, C-O and C-C bond, respectively [46,47,48]. The O 1s spectrum of V2O3/CD presented three contributions, with binding energies of 533.2, 532.0 and 531.3 eV, which were associated with O=C-O, C–OH and V-O, respectively [34,49]. In V 2p core-level spectrum, the two peaks located at 524.6 and 517.5 eV were caused by the spin–orbit splitting of V 2p1/2 and V 2p3/2, respectively, which is characteristic of vanadium in the +3 oxidation state [14,50]. Additionally, the energy difference between the binding energy of the O 1s and V 2p3/2 level (ΔE = EO 1s − E V 2p3/2) could be employed to determine the oxidation states of the vanadium oxides. The ΔE value of as-obtained V2O3@CD in the present work is 14 eV, which is consistent with the values of V3+ compounds reported in the literature, thus confirming the vanadium valence in +3 oxidation states [44]. Therefore, the XPS spectrum further illustrates the formation of V2O3 in the composites.
TGA analyses were conducted to determine the mass content of each component in the V2O3@CD composites. Figure 1g–i reveal the TGA curves of three V2O3@CD composites. Obviously, three weight-loss stages during the decomposition of the composites are presented in the TGA curves. The first and second stages of weight loss appear at about 130 °C and 400 °C, which are correlated with the release of adsorptive and structural water, respectively. The third stage of weight loss occurs at 380–570 °C, which is mainly associated with the combustion of carbon. The weight loss of VOCD-1, VOCD-2 and VOCD-3 in the third stage is 18.94%, 20.14% and 11.11%, respectively. Based on the results of TGA analysis, the mass contents of biomass carbon and vanadium oxide in VOCD-1, VOCD-2 and VOCD-3 are 18.94% and 69.71%, 20.14% and 73.37%, and 11.11% and 75.86%, respectively.
The morphology and structure of as-obtained V2O3@CD composites were investigated using SEM, TEM and EDS, and the results are presented in Figure 2. It can be discerned that the VOCD-1 composites show an obvious petal shape (Figure 2a,b), which may be due to the low content of V2O3 in the VOCD-1 composites. As displayed in Figure 3c,d, the structure of V2O3 in the VOCD-2 composites is mainly spherical with a diameter of 0.5–1.3 μm and a small number of sheets. The VOCD-3 composites display a unique porous architecture, and the diameter of pores is 6–12 μm (Figure 2e), indicating that the incorporation of V2O3 has little effect on the formation of the natural porous network of CD. As illustrated in Figure 2f,g, the structure of V2O3 in the VOCD-3 composites is spherical, with almost no sheets, and these microspheres are evenly filled in the macropores of CD. It is noteworthy that the surfaces of the V2O3 microspheres in the VOCD-2 and VOCD-3 composites are very coarse, indicating that both have a well-developed pore structure. The diffraction rings obtained from selected-area electron diffraction (SEAD) analysis (inset of Figure 2h) indicate the polycrystalline character of the sample. The inserted SAED pattern (Figure 2i) exhibits four distinctive diffraction rings belonging to the (012), (110), (113) and (116) planes of V2O3, respectively, demonstrating the existence of V2O3 in V2O3@CD. In addition, the HR-TEM image (inset of Figure 2i) manifests a crystal structure with clear lattice fringe of 2.47 Å that can be assigned to the d-spacing of (110) planes of the rhombohedral phase of crystalline V2O3, further verifying the presence of V2O3 in V2O3@CD [51]. The corresponding element mapping images (Figure 2j) reveal the homogenous distribution of C, O and V elements in the composites, indicating that V2O3 is homogeneously distributed on the surface and within the macropores of the porous CD matrix. The porous carbon with evenly distributed V2O3 is anticipated to endow V2O3@CD composites significantly increased pathways for alleviating the volume expansion, favoring the diffusion of electrolyte and improving the transfer of the electron, thus enhancing the electrochemical storage capacity of zinc.

3. Discussion

To explore the electrochemical behavior of V2O3@CD electrodes, EIS tests were carried out and corresponding Nyquist plots and the equivalent circuit are gathered in Figure 3a. The curves of V2O3@CD electrodes consist of one sloped line at low frequencies, associated with Warburg impedance and one semi-circle at a high-to-medium frequency region, which could be regarded as the charge transfer resistance (Rct) [52]. The VOCD-3 electrode exhibits the lowest Rct of 187.7 Ω compared to VOCD-1 (307.9 Ω) and VOCD-2 (220.4 Ω), demonstrating that the VOCD-3 electrode has the best electronic conductivity. Although V2O3 has higher resistance and ion resistance than CD, the Rct value is also related to the pore structure and the bonding strength between CD [53]. Therefore, the Rct of VOCD-3 is the lowest among the composites. The CV curves were administered to evaluate the kinetics of the electrochemical process of V2O3@CD electrodes, and the initial three cycles are illustrated in Figure 3b–d. The CV curves exhibit similar shapes, with two coupled redox peaks located at approximately 0.62/0.81 V and 0.91/1.11 V, manifesting that a multi-step reaction associated with Zn2+ ion (de)insertion process occurs in the V2O3@CD electrode [9,54]. In addition, VOCD-3 presents the widest redox peaks in the CV curve among the V2O3@CD composites, which is conducive to the kinetics of Zn2+ (de)insertion, thus exhibiting the highest specific capacity [55]. It is worth noting that with an increase in cycle times, the area in the CV curve also increases, illustrating an increase in the capacity of V2O3@CD electrodes at the early stage of the electrochemical cycle process (see Figure 3g).
The galvanostatic charge/discharge (GCD) profiles of V2O3@CD electrodes cycled at 50 mA g−1 between 0.2 and 1.6 V versus Zn2+/Zn are displayed in Figure 3e. Multiple redox pairs of voltage plateaus at 0.9/1.1 and 0.6/0.8 V can be observed on each charge and discharge curve, which can be consistent with the CV analysis. Remarkably, VOCD-3 can obtain a distinguished reversible capacity of 281.94 mAh g−1, which is higher than VOCD-1 of 264.47 mAh g−1 and VOCD-2 of 281.79 mAh g−1 at the initial cycle. The rate capabilities at different current densities, ranging from 0.05 C to 3 C for the V2O3@CD electrodes, are depicted in Figure 3f. In the three groups of rate capability, the specific capacity continues to decrease, regardless of the increase in current densities. VOCD-3 reveals significantly higher discharge capacities of 276.5, 231.60, 210.61, 193.60, 171.76, 155.81 and 151.38 mAh g−1 at 0.05, 0.1, 0.2, 0.5, 1, 2 and 3 C, respectively, compared to VOCD-1 and VOCD-2. When the current density is withdrawn to 0.05 C, VOCD-2 and VOCD-3 electrodes can recover to a high capacity, maintained at 257.33 and 270.24 mAh g−1, corresponding to 98.82% and 97.74% of their initial capacity, respectively, while VOCD-1 can retain only 69.41% of its initial capacity (184.86 mAh g−1) (Figure 3f,g). The above results demonstrate that the VOCD-3 electrode possesses the best rate performance compared with VOCD-1 and VOCD-2, which can be mainly attributed to the unique porous architectures of CD and high content of V2O3.
Figure 3h presents the long-cycle performance of V2O3@CD electrodes at 1 A g−1. The capacity of the three V2O3@CD composites increases gradually in the initial cycles, which is probably correlated with a gradual electrochemical activation process widely existing in vanadium-based electrodes of AZIBs [6,13,46]. Remarkably, VOCD-3 exhibits the highest capacity, up to 170.19 mAh g−1 after 693 cycles, among the three V2O3@CD composites. After 1000 cycles, the capacity of VOCD-3 slightly decreases to 151.89 mAh g−1, possessing capacity retention of 89.2%. In contrast, VOCD-1 and VOCD-2 show much lower capacities of 46.27 and 104.12 mAh g−1, with inferior capacity retentions of 38.5% and 77.2%, respectively, mainly due to their higher content of CD, resulting in lower reversible capacity. It has been reported that low levels of V2O3 may not take advantage of the benefits of V2O3, while high levels of V2O3 may lead to low electrical conductivity and poor CD protection [34]. The excellent electrochemical performance of VOCD-3 can be ascribed to the high content of V2O3 active and unique porous architectures in the CD matrix. The high content of V2O3 is beneficial to achieving a high reversible capacity. The porous architecture of CD not only effectively promotes the entrance of the electrolyte but can also perform as a buffering effect to assuage the huge volume change of V2O3 during the repeated (de)intercalation of Zn2+ from/into V2O3 process [15,34,56].
For a comparison, the electrochemical properties of V-based materials utilized as AZIB cathodes are collected in Table 1. The result shows that the V2O3/CD composites proposed in this work exhibit advantages. The excellent properties of V2O3/CD composites can probably be explained by the suitable ratio of V2O3 and CD to maximize their individual benefits.
To verify the structure stability, the morphology of the VOCD-3 electrode at diverse stages of pristine and after 300, 600 and 900 cycles were recorded by SEM, respectively, as shown in Figure 4a–d. In the initial state, nanoparticles are homogeneously distributed on the current collector without experiencing an agglomeration phenomenon in Figure 4a. After 300 cycles, the morphology of the VOCD-3 electrode remains almost the same as its original state, indicating good structural stability in the cycling process. After 900 cycles, apparent cracks emerged on the VOCD-3 electrode surface (Figure 4d), revealing that VOCD-3 is pulverizing during repetitive charge/discharge, thus resulting in capacity decline. From the element distribution mapping, in addition to the four elements of V, O, S and C, Zn is also detected from the electrode material, indicating that Zn2+ is successfully embedded in VOCD-3. In particular, no obvious dendrites are observed in the VOCD-3 electrode at the 200, 500 and 900 cycles (see Figure 4a–d). Combined with the favorable properties of Zn2+ storage, the protection mechanism of the porous structure can be intimated. Therefore, the as-obtained VOCD-3 composites employed as electrodes are conducive to improving the durability of the AZIBs.

4. Materials and Methods

4.1. Materials Synthesis

Urea (CH4N2O, ≥99.0%, AR), ethylene glycol (C2H6O2, AR) and ammonium metavanadate (NH4VO3, ≥99.0%, AR) were directly utilized without any purification. The V2O3@C composites were synthesized via evaporation-induced self-assembly technique. The detailed preparation process is schematically depicted in Figure 5. Firstly, 3 g CH4N2O and 2.2 g NH4VO3 were dissolved in a beaker with 20 mL of distilled water. After continuous magnetic stirring at 60 °C for 0.5 h, 20 mL C2H6O2 and 0.4 g dried dictyophora were added into the above-mentioned beaker and stirred for 0.5 h to form homogeneous liquid. Then, the obtained solution was heated completely in an oven at 100 °C for 10 h. The resultant powder sample was sintered first at 350 °C for 4 h in an argon-filled tube furnace to obtain VO2 and carbonized dictyophora. Then, the obtained products were calcined at 800 °C for 8 h, with a heating rate of 10 °C min−1. The chemical reactions occurring in this process and the reaction principle can be described by Gibbs free energy (G°) as a function of temperature (T), as presented in Equation (1).
2 VO 2 + C V 2 O 3 + CO ,   Δ G °   J / mol = 95300 158.68 T
According to the thermodynamic data, T = (t +273.5) K = 1073.5 K, ∆G° = 95300 − 158.68 T < 0, indicating that the reaction is positive. In this reaction, VO2 can be deoxidized to V2O3 using carbon as the reducing agent. The C acts as a reducing agent to reduce the high-valent VO2 to low-valent V2O3.
The final synthesized sample was named VOCD-1. For comparison, VOCD-2 and VOCD-3 were also prepared in the same way by changing the mass ratio of dictyophora to NH4VO3. The dosage of raw materials for each sample is summarized in Table 2.

4.2. Materials Characterization

The crystal structure of the synthesized samples was characterized by a powder Bruker D8 Advance X-ray diffractometer (XRD, Rigaku Corporation, Japan)) equipped with Cu Kα radiation (λ = 0.152 nm). The FT-IR spectra with a wavenumber range of 400–4000 cm−1 were recorded on an FT-IR spectrometer (Vertex-70, Bruker, Heidelberg, Germany). X-ray photoelectron spectroscopy (XPS, Kratos Analytical Ltd., Manchester, UK) analysis was conducted on a Thermo Escalab 250Xi spectrometer (Thermo Fisher Scientific, Waltham, MA, USA) employing a monochromatic Al-Kα X-ray source (Kα = 1486.6 eV). The measured binding energies were calibrated by using the containment carbon (C1s = 284.6 eV) as a reference. All XPS spectra analysis and curve fitting were carried out using PeakFit v4.12 software. The surface morphology and microstructure were determined by transmission electron microscopy equipped with Energy dispersive spectra (TEM-EDS, Tecnai F20, FEI Inc., Valley City, ND, USA) and scanning electron microscopic (SEM, Quanta 200S, FEI, Inc., Valley City, ND, USA) techniques. Thermogravimetric (TG) analyzer (TG, Netzsch, STA409C, Frankfurt, Germany) was applied to evaluate the thermal stability of the samples under nitrogen atmosphere.

4.3. Electrochemical Measurements

The cathode was composed of active materials, activated carbon and polyvinylidene fluoride (PVDF), with a mass ratio of 7:2:1, through mixing N-methyl-2-pyrrolidone (NMP) solvent to form slurry. After that, the slurry was pasted uniformly on a 316 L stainless-steel foil, subsequently heated at 100 °C overnight in oven and cut into Φ14 mm discs. The ZnSO4 solution (3 mol∙L−1) was applied as electrolyte, and zinc foil and glass-fiber film were applied as counter electrode and separator, respectively. The CR2025 coin batteries were assembled under normal air atmosphere conditions. The batteries were galvanostatically measured with various charge and discharge rates through an NEWARE testing instrument. The CHI660E electrochemical workstation is utilized to test the cyclic voltammetry (CV) and electrochemical impedance spectroscopy (EIS).

5. Conclusions

In summary, a facile and cost-effective V2O3@CD composite as a cathode for AZIBs was successfully fabricated via an evaporation-induced self-assembly process. The resultant composite reveals a unique structure in which V2O3 is homogeneously embedded in the macropores of bio-carbon with porous architectures. When V2O3@CD composites are employed as AZIB cathodes, the carbon skeleton can effectively enhance the conductivity and restrain the large volume change, guaranteeing the structural stability of the composites. As a consequence, the VOCD-3 composites exhibit the lowest Rct of 187.7 Ω as well as a superior rate capability of 151 mAh g−1 at 3 A g−1 among the as-prepared V2O3@CD composites. In addition, the maximum capacity of 270 mAh g−1 at 1 A g−1 after 693 cycles is obtained in the VOCD-3 composites. It is believed that this study on a V2O3/biomass-derived carbon electrode could provide a new approach for the development of high-performance AZIBs using low-cost biomass as raw materials.

Author Contributions

Conceptualization and methodology, W.Z. and G.Z.; software, M.H., H.J. and G.Z.; validation, W.Z., G.Z. and H.C.; formal analysis, G.Z. and H.J.; data curation, G.Z. and M.H.; writing—original draft preparation, G.Z.; writing—review and editing, W.Z. and C.Z.; supervision and project administration, W.Z., C.Z. and S.J.; funding acquisition, W.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Natural Science Foundation of Hunan Province (No. 2022JJ50086).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors appreciate the support of the College of Metallurgy and Materials Engineering Hunan University of Technology, Nanjing Forestry University, and Changsha University.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not applicable.

References

  1. Pang, Q.; Sun, C.; Yu, Y.; Zhao, K.; Zhang, Z.; Voyles, P.M.; Chen, G.; Wei, Y.; Wang, X. H2V3O8 Nanowire/Graphene Electrodes for Aqueous Rechargeable Zinc Ion Batteries with High Rate Capability and Large Capacity. Adv. Energy Mater. 2018, 8, 1800144. [Google Scholar] [CrossRef]
  2. Tamilselvan, M.; Sreekanth, T.; Yoo, K.; Kim, J. Binder-Free Coaxially Grown V6O13 Nanobelts on Carbon Cloth as Cathodes for Highly Reversible Aqueous Zinc Ion Batteries. Appl. Surf. Sci. 2020, 529, 147077. [Google Scholar] [CrossRef]
  3. Yu, J.; Liu, S.; Duan, G.; Fang, H.; Hou, H. Dense and thin coating of gel polymer electrolyte on sulfur cathode toward high performance Li-sulfur battery. Compos. Commun. 2020, 19, 239–245. [Google Scholar] [CrossRef]
  4. Chen, S.; Jeong, S.R.; Tao, S. Key materials and future perspective for aqueous rechargeable lithium-ion batteries. Mater. Rep. Energy 2022, 2, 100096. [Google Scholar] [CrossRef]
  5. Zhang, H.; Yao, Z.; Lan, D.; Liu, Y.; Cui, J. N-doped carbon/V2O3 microfibers as high-rate and ultralong-life cathode for rechargeable aqueous zinc-ion batteries. J. Alloys Compd. 2020, 861, 158560. [Google Scholar] [CrossRef]
  6. Liu, Y.Y.; Lv, T.T.; Wang, H.; Guo, X.T.; Pang, H. Nsutite-Type VO2 Microcrystals as Highly Durable Cathode Materials for Aqueous Zinc-Ion Batteries. Chem. Eng. J. 2021, 417, 128408. [Google Scholar] [CrossRef]
  7. Chen, X.; Hu, X.; Chen, Y.; Cao, X.; Huang, Y.; Zhang, H.; Liu, J.-H.; Wang, Y.; Chou, S.-L.; Cao, D. Ultrastable hydrated vanadium dioxide cathodes for high-performance aqueous zinc ion batteries with H+/Zn2+ Co-insertion mechanism. J. Mater. Chem. A 2022, 10, 22194–22204. [Google Scholar] [CrossRef]
  8. Zhang, Y.F.; Xu, L.; Jiang, H.M.; Liu, Y.Y.; Meng, C.G. Polyaniline-expanded the interlayer spacing of hydrated vanadium pentoxide by the interface-intercalation for aqueous rechargeable Zn-ion batteries. J. Colloid Interface Sci. 2021, 603, 641–650. [Google Scholar] [CrossRef] [PubMed]
  9. Chen, H.; Rong, Y.; Yang, Z.; Deng, L.; Wu, J. V2O3@Amorphous Carbon as a Cathode of Zinc Ion Batteries with High Stability and Long Cycling Life. Ind. Eng. Chem. Res. 2021, 60, 1517–1525. [Google Scholar] [CrossRef]
  10. Jiang, H.; Zhang, Y.; Waqar, M.; Yang, J.; Liu, Y.; Sun, J.; Feng, Z.; Sun, J.; Pan, Z.; Meng, C.; et al. Anomalous Zn2+ Storage Behavior in Dual-Ion-In-Sequence Reconstructed Vanadium Oxides. Adv. Funct. Mater. 2023, 33, 2213127. [Google Scholar] [CrossRef]
  11. Li, J.; Xue, H.; Xu, N.; Zhang, X.; Wang, Y.; He, R.; Huang, H.; Qiao, J. Co/Ni dual-metal embedded in heteroatom doped porous carbon core-shell bifunctional electrocatalyst for rechargeable Zn-air batteries. Mater. Rep. Energy 2022, 2, 100090. [Google Scholar] [CrossRef]
  12. Guo, W.; Guo, X.; Yang, L.; Wang, T.; Zhang, M.; Duan, G.; Liu, X.; Li, Y. Synthetic melanin facilitates MnO supercapacitors with high specific capacitance and wide operation potential window. Polymer 2021, 235, 124276. [Google Scholar] [CrossRef]
  13. Ding, Y.C.; Peng, Y.Q.; Chen, S.H.; Zhang, X.X.; Li, Z.Q.; Zhu, L.; Mo, L.E.; Hu, L.H. Hierarchical Porous Metallic V2O3@C for Advanced Aqueous Zinc-Ion Batteries. Acs Appl. Mater. Interfaces 2019, 11, 44109–44117. [Google Scholar] [CrossRef] [PubMed]
  14. Zhang, H.; Zhang, Y.; Liu, Y.; Shi, X.; Zhang, Y.; Bai, L.; Wang, Q.; Sun, L. Oxygen-Deficient α-MnO2 Nanotube/Graphene/N, P Codoped Porous Carbon Composite Cathode To Achieve High-Performing Zinc-Ion Batteries. ACS Appl. Mater. Interfaces 2022, 14, 36668–36678. [Google Scholar] [CrossRef] [PubMed]
  15. Gao, Q.-L.; Li, D.-S.; Liu, X.-M.; Wang, Y.-F.; Liu, W.-L.; Ren, M.-M.; Kong, F.; Wang, S.-J.; Zhou, R.-C. Biomass-derived mesoporous carbons materials coated by α-Mn3O4 with ultrafast zinc-ion diffusion ability as cathode for aqueous zinc ion batteries. Electrochim. Acta 2020, 335, 135642. [Google Scholar] [CrossRef]
  16. Zhu, X.D.; Cao, Z.Y.; Li, X.L.; Pei, L.Y.; Jones, J.; Zhou, Y.N.; Dong, P.; Wang, L.P.; Xe, M.X.; Shen, J.F. Ion-intercalation regulation of MXene-derived hydrated vanadates for high-rate and long-life Zn-Ion batteries. Energy Storage Mater. 2022, 45, 568–577. [Google Scholar] [CrossRef]
  17. Lv, T.T.; Luo, X.; Yuan, G.Q.; Yang, S.Y.; Pang, H. Layered VO2@N-doped carbon composites for high-performance rechargeable aqueous zinc-ion batteries. Chem. Eng. J. 2022, 428, 131211. [Google Scholar] [CrossRef]
  18. Niu, Y.; Wang, D.H.; Ma, Y.J.; Zhi, L.J. Cascading V2O3/N-doped carbon hybrid nanosheets as high-performance cathode materials for aqueous zinc-ion batteries. Chin. Chem. Lett. 2022, 33, 1430–1434. [Google Scholar] [CrossRef]
  19. Zheng, J.; Zhan, C.Y.; Zhang, K.; Fu, W.W.; Nie, Q.J.; Zhang, M.; Shen, Z.R. Rapid Electrochemical Activation of V2O3@C Cathode for High-Performance Zinc-Ion Batteries in Water-in-Salt Electrolyte. Chemsuschem 2022, 15, e202200075. [Google Scholar] [CrossRef]
  20. Ding, J.W.; Zheng, H.Y.; Gao, H.G.; Liu, Q.N.; Hu, Z.; Han, L.F.; Wang, S.W.; Wu, S.D.; Fang, S.M.; Chou, S.L. In Situ Lattice Tunnel Distortion of Vanadium Trioxide for Enhancing Zinc Ion Storage. Adv. Energy Mater. 2021, 11, 2100973. [Google Scholar] [CrossRef]
  21. Du, Y.H.; Wang, X.Y.; Zhang, Y.; Zhang, H.B.; Man, J.Z.; Liu, K.; Sun, J.C. High mass loading CaV4O9 microflowers with amorphous phase transformation as cathode for aqueous zinc-ion battery. Chem. Eng. J. 2022, 434, 134642. [Google Scholar] [CrossRef]
  22. Chen, L.N.; Ruan, Y.S.; Zhang, G.B.; Wei, Q.L.; Jiang, Y.L.; Xiong, T.F.; He, P.; Yang, W.; Yan, M.Y.; An, Q.Y.; et al. Ultrastable and High-Performance Zn/VO2 Battery Based on a Reversible Single-Phase Reaction (vol 31, pg 699, 2019). Chem. Mater. 2019, 31, 5342. [Google Scholar] [CrossRef] [Green Version]
  23. Kulkarni, P.; Beere, H.K.; Jalalah, M.; Alsaiari, M.; Balakrishna, R.G.; Harraz, F.A.; Ghosh, D. Developing a high-performance aqueous zinc battery with Zn2+ pre-intercalated V3O7middotH2O cathode coupled with surface engineered metallic zinc anode. J. Electroanal. Chem. 2022, 924, 116851. [Google Scholar] [CrossRef]
  24. Zhu, Y.J.; Lin, X.R.; Jin, X.; Han, T.L.; Zhang, H.G.; Liu, J.Y. A flexible self-healing Zn3V2O7(OH)(2)center dot 2H(2)O-based Zn-ion battery under continuous folding and twisting. Chem. Commun. 2022, 58, 8117–8120. [Google Scholar] [CrossRef] [PubMed]
  25. Chen, Y.; Zhang, Q.; Chi, M.; Guo, C.; Wang, S.; Min, D. Preparation and performance of different carbonized wood electrodes. J. For. Eng. 2022, 7, 127–135. [Google Scholar]
  26. Wang, R.; Xuelian, Z.; Xu, T.; Bian, H.; Dai, H. Research progress on the preparation of lignin-derived carbon dots and graphene quantum dots. J. For. Eng. 2021, 6, 29–37. [Google Scholar]
  27. Hu, J.X.; Xie, Y.Y.; Zheng, J.Q.; Li, H.Z.; Wang, T.S.; Lai, Y.Q.; Zhang, Z.A. Encapsulating V2O3 Nanoparticles in Hierarchical Porous Carbon Nanosheets via C-O-V Bonds for Fast and Durable Potassium-Ion Storage. Acs Appl. Mater. Interfaces 2021, 13, 12149–12158. [Google Scholar] [CrossRef]
  28. Du, Y.H.; Wang, X.Y.; Man, J.Z.; Sun, J.C. A novel organic-inorganic hybrid V2O5@polyaniline as high-performance cathode for aqueous zinc-ion batteries. Mater. Lett. 2020, 272, 127813. [Google Scholar] [CrossRef]
  29. Jiang, L.; Qu, Y.; Ren, Z.Y.; Yu, P.; Zhao, D.D.; Zhou, W.; Wang, L.; Fu, H.G. In Situ Carbon-Coated Yolk-Shell V2O3 Microspheres for Lithium-Ion Batteries. Acs Appl. Mater. Interfaces 2015, 7, 1595–1601. [Google Scholar] [CrossRef]
  30. Zheng, S.; Zhang, J.; Deng, H.; Du, Y.; Shi, X. Chitin derived nitrogen-doped porous carbons with ultrahigh specific surface area and tailored hierarchical porosity for high performance supercapacitors. J. Bioresour. Bioprod. 2021, 6, 142–151. [Google Scholar] [CrossRef]
  31. Yan, B.; Zheng, J.; Wang, F.; Zhao, L.; Zhang, Q.; Xu, W.; He, S. Review on porous carbon materials engineered by ZnO templates: Design, synthesis and capacitance performance. Mater. Des. 2021, 201, 109518. [Google Scholar] [CrossRef]
  32. He, Z.X.; Cheng, G.; Jiang, Y.Q.; Li, Y.H.; Zhu, J.; Meng, W.; Zhou, H.Z.; Dai, L.; Wang, L. Novel 2D porous carbon nanosheet derived from biomass: Ultrahigh porosity and excellent performances toward V2+/V3+ redox reaction for vanadium redox flow battery. Int. J. Hydrogen Energy 2020, 45, 3959–3970. [Google Scholar] [CrossRef]
  33. Li, R.; Lou, Z.; Gu, S.; Wang, Q.; Liu, J.; Yanjun, L. Preparation of magnetic carbon with microwave absorption property using bamboo powder. J. For. Eng. 2021, 6, 112–120. [Google Scholar]
  34. Li, Y.; Lin, W.; Xue, L.; Xie, J.; Wei, B.; Chen, G.; Chen, D. Facile preparation of V2O3/black fungus-derived carbon composite with hierarchical porosity as a promising electrode for lithium/sodium ion batteries. J. Alloys Compd. 2022, 905, 164258. [Google Scholar] [CrossRef]
  35. Xiao, J.; Li, H.; Zhang, H.; He, S.; Zhang, Q.; Liu, K.; Jiang, S.; Duan, G.; Zhang, K. Nanocellulose and its derived composite electrodes toward supercapacitors: Fabrication, properties, and challenges. J. Bioresour. Bioprod. 2022, 7, 245–269. [Google Scholar] [CrossRef]
  36. Duan, G.; Zhao, L.; Chen, L.; Wang, F.; He, S.; Jiang, S.; Zhang, Q. ZnCl2 regulated flax-based porous carbon fibers for supercapacitors with good cycling stability. New J. Chem. 2021, 45, 22602–22609. [Google Scholar] [CrossRef]
  37. Zhang, T.; Zhang, L.; Zhao, L.N.; Huang, X.X.; Li, W.; Li, T.; Shen, T.; Sun, S.N.; Hou, Y.L. Free-Standing, Foldable V2O3/Multichannel Carbon Nanofibers Electrode for Flexible Li-Ion Batteries with Ultralong Lifespan. Small 2020, 16, 2005302. [Google Scholar] [CrossRef]
  38. Zakharova, G.S.; Thauer, E.; Enyashin, A.N.; Deeg, L.F.; Zhu, Q.; Klingeler, R. V2O3/C composite fabricated by carboxylic acid-assisted sol-gel synthesis as anode material for lithium-ion batteries. J. Sol-Gel Sci. Technol. 2021, 98, 549–558. [Google Scholar] [CrossRef]
  39. Bai, Y.C.; Tang, Y.K.; Liu, L.; Li, X.H.; Gao, Y. Peapod-like CNT@V2O3 with Superior Electrochemical Performance as an Anode for Lithium-Ion Batteries. Acs Sustain. Chem. Eng. 2018, 6, 14614–14620. [Google Scholar] [CrossRef]
  40. Feng, Z.Y.; Zhang, Y.F.; Yu, X.M.; Yu, Y.T.; Huang, C.; Meng, C.G. Aluminum-ion intercalation and reduced graphene oxide wrapping enable the electrochemical properties of hydrated V2O5 for Zn-ion storage. Colloids Surf. A-Physicochem. Eng. Asp. 2022, 641, 128473. [Google Scholar] [CrossRef]
  41. Liu, Y.Y.; Pan, Z.H.; Tian, D.; Hu, T.; Jiang, H.M.; Yang, J.; Sun, J.J.; Zheng, J.Q.; Meng, C.G.; Zhang, Y.F. Employing “one for two” strategy to design polyaniline-intercalated hydrated vanadium oxide with expanded interlayer spacing for high-performance aqueous zinc-ion batteries. Chem. Eng. J. 2020, 399, 125842. [Google Scholar] [CrossRef]
  42. Kim, J.H.; Kim, Y.S.; Moon, S.H.; Park, D.H.; Kim, M.C.; Choi, J.H.; Shin, J.H.; Park, K.W. Enhanced electrochemical performance of a selectively formed V2O3/C composite structure for Li-ion batteries. Electrochim. Acta 2021, 389, 138685. [Google Scholar] [CrossRef]
  43. Henry, A.; Hesemann, P.; Alauzun, J.G.; Boury, B. Reductive mineralization of cellulose with vanadium, iron and tungsten chlorides and access to MxOy metal oxides and MxOy/C metal oxide/carbon composites. Carbohydr. Polym. 2017, 174, 697–705. [Google Scholar] [CrossRef] [PubMed]
  44. Zhang, Y.F.; Fan, M.J.; Liu, X.H.; Huang, C.; Li, H.B. Beltlike V2O3@C Core-Shell-Structured Composite: Design, Preparation, Characterization, Phase Transition, and Improvement of Electrochemical Properties of V2O3. Eur. J. Inorg. Chem. 2012, 10, 1650–1659. [Google Scholar] [CrossRef]
  45. Zheng, J.Q.; Zhang, Y.F.; Jing, X.Y.; Liu, X.Y.; Hu, T.; Lv, T.M.; Zhang, S.Q.; Meng, C.G. Synthesis of amorphous carbon coated on V2O3 core-shell composites for enhancing the electrochemical properties of V2O3 as supercapacitor electrode. Colloids Surf. A-Physicochem. Eng. Asp. 2017, 518, 188–196. [Google Scholar] [CrossRef]
  46. Liu, X.D.; Wang, Z.Q.; Niu, Y.L.; Liu, C.Y.; Chen, H.M.; Ren, X.Z.; Wang, M.S.; Lau, W.M.; Zhou, D. Scalable synthesis of novel V2O3/carbon composite as advanced cathode material for aqueous zinc-ion batteries. Ceram. Int. 2022, 48, 15594–15602. [Google Scholar] [CrossRef]
  47. Cui, T.; Wang, Y.P.; Ye, T.; Wu, J.; Chen, Z.; Li, J.; Lei, Y.; Wang, D.; Li, Y. Engineering Dual Single-Atom Sites on 2D Ultrathin N-doped Carbon Nanosheets Attaining Ultra-Low-Temperature Zinc-Air Battery. Angew. Chem. Int. Ed. Engl. 2022, 61, e202115219. [Google Scholar] [CrossRef]
  48. Wang, Y.; Xu, L.; Zhan, L.; Yang, P.; Tang, S.; Liu, M.; Zhao, X.; Xiong, Y.; Chen, Z.; Lei, Y. Electron accumulation enables Bi efficient CO2 reduction for formate production to boost clean Zn-CO2 batteries. Nano Energy 2022, 92, 106780. [Google Scholar] [CrossRef]
  49. Wang, Q.; Feng, Q.; Lei, Y.; Tang, S.; Xu, L.; Xiong, Y.; Fang, G.; Wang, Y.; Yang, P.; Liu, J.; et al. Quasi-solid-state Zn-air batteries with an atomically dispersed cobalt electrocatalyst and organohydrogel electrolyte. Nat. Commun. 2022, 13, 3689. [Google Scholar] [CrossRef]
  50. Zhao, J.; Zhao, Y.Y.; Yue, W.C.; Li, X.; Gao, N.; Zhang, Y.J.; Hu, C.Q. V2O3/VN Catalysts Decorated Free-Standing Multifunctional Interlayer for High-Performance Li-S Battery. Chem. Eng. J. 2022, 441, 136082. [Google Scholar] [CrossRef]
  51. Saroha, R.; Heo, J.; Liu, Y.; Angulakshmi, N.; Lee, Y.; Cho, K.-K.; Ahn, H.-J.; Ahn, J.-H. V2O3-decorated carbon nanofibers as a robust interlayer for long-lived, high-performance, room-temperature sodium–sulfur batteries. Chem. Eng. J. 2022, 431, 134205. [Google Scholar] [CrossRef]
  52. Won, J.M.; Ko, Y.N.; Lee, J.K.; Kang, Y.C. Superior electrochemical properties of rutile VO2-carbon composite microspheres as a promising anode material for lithium ion batteries. Electrochim. Acta 2015, 156, 179–187. [Google Scholar] [CrossRef]
  53. Xiao, B.; Zhang, B.; Tang, L.B.; An, C.S.; He, Z.J.; Tong, H.; Yu, W.J.; Zheng, J.C. V2O3/rGO composite as a potential anode material for lithium ion batteries. Ceram. Int. 2018, 44, 15044–15049. [Google Scholar] [CrossRef]
  54. Liu, Y.; Liu, Y.; Wu, X.; Cho, Y.R. Enhanced Electrochemical Performance of Zn/VOx Batteries by a Carbon-Encapsulation Strategy. Acs Appl. Mater. Interfaces 2022, 14, 11654–11662. [Google Scholar] [CrossRef]
  55. Ren, H.-Z.; Zhang, J.; Wang, B.; Luo, H.; Jin, F.; Zhang, T.-R.; Ding, A.; Cong, B.-W.; Wang, D.-L. A V2O3@N–C cathode material for aqueous zinc-ion batteries with boosted zinc-ion storage performance. Rare Met. 2022, 41, 1605–1615. [Google Scholar] [CrossRef]
  56. Dai, X.; Wan, F.; Zhang, L.L.; Cao, H.M.; Niu, Z.Q. Freestanding graphene/VO2 composite films for highly stable aqueous Zn-ion batteries with superior rate performance. Energy Storage Mater. 2019, 17, 143–150. [Google Scholar] [CrossRef]
  57. Deng, L.; Chen, H.; Wu, J.; Yang, Z.; Rong, Y.; Fu, Z. V2O3 as cathode of zinc ion battery with high stability and long cycling life. Ionics 2021, 27, 3393–3402. [Google Scholar] [CrossRef]
  58. Bin, D.; Wang, Y.; Tamirat, A.G.; Zhu, P.; Yang, B.; Wang, J.; Huang, J.; Xia, Y. Stable High-Voltage Aqueous Zinc Battery Based on Carbon-Coated NaVPO4F Cathode. ACS Sustain. Chem. Eng. 2021, 9, 3223–3231. [Google Scholar] [CrossRef]
  59. Venkatkarthick, R.; Rodthongkum, N.; Zhang, X.; Wang, S.; Pattananuwat, P.; Zhao, Y.; Liu, R.; Qin, J. Vanadium-Based Oxide on Two-Dimensional Vanadium Carbide MXene (V2Ox@V2CTx) as Cathode for Rechargeable Aqueous Zinc-Ion Batteries. ACS Appl. Energy Mater. 2020, 3, 4677–4689. [Google Scholar] [CrossRef]
  60. Li, Y.; Xu, P.; Jiang, J.; Yao, J.; Huang, B.; Yang, J. Facile synthesis of ultra-large V2O5 xerogel flakes and its application as a cathode material for aqueous Zn-ion batteries. Mater. Today Commun. 2021, 26, 101849. [Google Scholar] [CrossRef]
  61. Liu, X.; Wang, Z.; Niu, Y.; Liu, C.; Chen, H.; Ren, X.; Liu, Z.; Lau, W.-M.; Zhou, D. Electrospun V2O3@Carbon Nanofibers as a Flexible and Binder-Free Cathode for Highly Stable Aqueous Zn-Ion Full Batteries. ACS Appl. Energy Mater. 2022, 5, 3525–3535. [Google Scholar] [CrossRef]
  62. Lan, B.; Tang, C.; Chen, L.; Zhang, W.; Tang, W.; Zuo, C.; Fu, X.; Dong, S.; An, Q.; Luo, P. FeVO4⋅nH2O@rGO nanocomposite as high performance cathode materials for aqueous Zn-ion batteries. J. Alloys Compd. 2020, 818, 153372. [Google Scholar] [CrossRef]
Figure 1. (a) XRD patterns, (b) FT-IR spectrum, (c) XPS survey spectra of the V2O3/CD composites, high-resolution (d) C 1s, (e) O 1s and (f) V 2p XPS spectrum of the V2O3/CD composites and (gi) TGA curves of the V2O3/CD composites.
Figure 1. (a) XRD patterns, (b) FT-IR spectrum, (c) XPS survey spectra of the V2O3/CD composites, high-resolution (d) C 1s, (e) O 1s and (f) V 2p XPS spectrum of the V2O3/CD composites and (gi) TGA curves of the V2O3/CD composites.
Molecules 28 02147 g001
Figure 2. SEM images of (a,b) VOCD-1, (c,d) VOCD-2 and (eg) VOCD-3, (h,i) HR-TEM image (inset: SEAD pattern) and (j) elemental mapping images of the synthesized V2O3@CD.
Figure 2. SEM images of (a,b) VOCD-1, (c,d) VOCD-2 and (eg) VOCD-3, (h,i) HR-TEM image (inset: SEAD pattern) and (j) elemental mapping images of the synthesized V2O3@CD.
Molecules 28 02147 g002
Figure 3. (a) Nyquist plots of V2O3@CD electrodes (inset: equivalent circuit simulation). (bd) CV curves of the initial three cycles for the VOCD-2 and VOCD-3 within 0.2–2.0 V at 0.1 mV s−1. (e) Discharge–charge profiles at 0.05 A g−1. (f) Rate capability at various current densities. (g) Capacity retention after 1000 cycles at 1 A g−1, capacity retention after rate cycling to 3 A g−1 and rate cycling back to 0.05 A g−1. (h) Galvanostatic cycling at 1 A g−1.
Figure 3. (a) Nyquist plots of V2O3@CD electrodes (inset: equivalent circuit simulation). (bd) CV curves of the initial three cycles for the VOCD-2 and VOCD-3 within 0.2–2.0 V at 0.1 mV s−1. (e) Discharge–charge profiles at 0.05 A g−1. (f) Rate capability at various current densities. (g) Capacity retention after 1000 cycles at 1 A g−1, capacity retention after rate cycling to 3 A g−1 and rate cycling back to 0.05 A g−1. (h) Galvanostatic cycling at 1 A g−1.
Molecules 28 02147 g003
Figure 4. SEM images of VOCD-3 electrodes at different states of (a) pristine, after (b) 300, (c) 600 and (d) 900 cycles; (e) EDS mapping image of the V, O, S, C and Zn elemental distribution in the VOCD-3 cathode after 900 cycles.
Figure 4. SEM images of VOCD-3 electrodes at different states of (a) pristine, after (b) 300, (c) 600 and (d) 900 cycles; (e) EDS mapping image of the V, O, S, C and Zn elemental distribution in the VOCD-3 cathode after 900 cycles.
Molecules 28 02147 g004
Figure 5. The preparation process of V2O3@CD composites.
Figure 5. The preparation process of V2O3@CD composites.
Molecules 28 02147 g005
Table 1. Comparison of the electrochemical performance of V2O3/CD composites in this work with other previously reported V-based cathode materials for AZIBs.
Table 1. Comparison of the electrochemical performance of V2O3/CD composites in this work with other previously reported V-based cathode materials for AZIBs.
MaterialsMethod of SynthesisCurrent Density
(A g−1)
Specific Capacity
(mAh g−1)
Capacity RetentionCycle NumberRef./Year
V2O3@carbonized dictyophoraevaporation-induced self-assembly technique1151.989.24%1000This work
V2O3@amorphous carbonCalcination111690.7%1600[9]/2021
V2O3Reduction method of boron0.116176.9%100[57]/2021
Carbon-coated NaVPO4FCVD0.187.494.5%400[58]/2021
V2Ox@V2CTxHigh-temperature etching and electrochemical active187.381.6%200[59]/2020
V2O5 xerogel flakesHydrothermal113564%200[60]/2021
V2O3@Carbon NanofibersElectrospinning0.212080%1000[61]/2022
FeVO4•nH2O@rGOHydrothermal19243.8% a1000[62]/2020
a Values are estimated from the graphs.
Table 2. The dosage of each sample is summarized as follows.
Table 2. The dosage of each sample is summarized as follows.
SamplesDictyophoraC2H6O2CH4N2ONH4VO3
VOCD-10.40 g20 mL0.75 g0.55 g
VOCD-20.40 g20 mL1.50 g1.10 g
VOCD-30.40 g20 mL3.00 g2.20 g
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhou, W.; Zeng, G.; Jin, H.; Jiang, S.; Huang, M.; Zhang, C.; Chen, H. Bio-Template Synthesis of V2O3@Carbonized Dictyophora Composites for Advanced Aqueous Zinc-Ion Batteries. Molecules 2023, 28, 2147. https://doi.org/10.3390/molecules28052147

AMA Style

Zhou W, Zeng G, Jin H, Jiang S, Huang M, Zhang C, Chen H. Bio-Template Synthesis of V2O3@Carbonized Dictyophora Composites for Advanced Aqueous Zinc-Ion Batteries. Molecules. 2023; 28(5):2147. https://doi.org/10.3390/molecules28052147

Chicago/Turabian Style

Zhou, Wei, Guilin Zeng, Haotian Jin, Shaohua Jiang, Minjie Huang, Chunmei Zhang, and Han Chen. 2023. "Bio-Template Synthesis of V2O3@Carbonized Dictyophora Composites for Advanced Aqueous Zinc-Ion Batteries" Molecules 28, no. 5: 2147. https://doi.org/10.3390/molecules28052147

Article Metrics

Back to TopTop