Next Article in Journal
Effect of Lithium Salt Concentration on Materials Characteristics and Electrochemical Performance of Hybrid Inorganic/Polymer Solid Electrolyte for Solid-State Lithium-Ion Batteries
Next Article in Special Issue
Tuning the Architecture of Hierarchical Porous CoNiO2 Nanosheet for Enhanced Performance of Li-S Batteries
Previous Article in Journal
Li-Ion Battery Short-Circuit Protection by Voltage-Driven Switchable Resistance Polymer Layer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Microstructure Modulation of Zn Doped VO2(B) Nanorods with Improved Electrochemical Properties towards High Performance Aqueous Batteries

Laboratory of Magnetoelectronic Information, School of Physics and Electronic Engineering, Henan Key Functional Materials, Zhengzhou University of Light Industry, Zhengzhou 450002, China
*
Author to whom correspondence should be addressed.
Batteries 2022, 8(10), 172; https://doi.org/10.3390/batteries8100172
Submission received: 6 September 2022 / Revised: 27 September 2022 / Accepted: 29 September 2022 / Published: 9 October 2022
(This article belongs to the Special Issue Materials and Interface Designs for Batteries)

Abstract

:
Vanadium dioxide with monoclinic structure is theoretically a promising layered cathode material for aqueous metal-ion batteries due to its excellent specific capacity. However, its poor cycling stability limits its application as an electrode material. In this study, a series of Zn-doped VO2 (V1−xZnxO2) nanorods were successfully fabricated by the technology of one-step hydrothermal synthesis. The XRD result indicated that there was a slight lattice distortion caused by doped Zn2+ with a larger ion radius. The positron lifetime spectrum showed that there were vacancy cluster defects in all the samples. The electrochemical measurement demonstrated the enhancement of the specific capacitance of V1−xZnxO2 electrodes compared with the undoped sample. In addition, the discharge capacitance of the sample remained around 86% after 1000 charge/discharge cycles. This work proves that Zn2+ doping is a valid tactic for the application of nano-VO2(B) in energy storage electrode materials.

Graphical Abstract

1. Introduction

The rising demands for batteries in electric vehicles are driving the development of new energy conversion and storage technologies based on new materials with better overall electrochemical properties [1,2,3]. The next-generation aqueous metal-ion batteries, of high safety, low cost and high theoretical specific capacitance, are considered as promising candidates for the future energy storage applications [4,5]. The cathode materials, as the most important component affecting the performance of aqueous metal-ion batteries, remain the biggest bottleneck restricting their further practical applications.
The vanadium-based compounds have been widely used as electrode materials because of their high specific capacity and fast ion diffusion kinetics for aqueous metal-ion batteries, which arise from various chemical valences of Vx+(x = 3, 4, 5) and large interlayer spacing [6,7,8]. Particularly, monoclinic phase VO2(B) consists of two identical angular VO6 octahedrons superimposed along the B-axis, and it has a tunnel structure which is beneficial to metal ion embedding and de-embedding. In addition, the edge-sharing VO6 octahedron endows the metastable monoclinic VO2(B) with layered structural flexibility alleviating severe structural distortion during the extraction and insertion of metal ions in electrolytes during the electrochemical reaction process. It has been reported that the V2O5·xH2O xerogel cathode exhibits an initial capacity of 308 mAh·g−1 at 1.0–4.0 V [9]. A reversible capacity of about 200 mAh·g−1 can be achieved at 50 mA·g−1 in metal-ion batteries between 1.5 and 4.0 V [10]. However, the VO2(B) electrode materials still face a severe capacitance fading problem in practical applications [11]. This may be related to the interface reaction between the vanadium and electrolyte in the process of acid–base interaction, which aggravates the dissolution problem of VO2(B) [12]. As a result, the loss of active materials can trigger the fast capacity fade caused by the catalytic decomposition of the electrolyte. This problem has been a serious obstacle for the development of VO2(B) electrode materials.
The electrode materials are the critical factors determining the cyclability of aqueous metal-ion batteries, and some of the literature has reported that doping with metal cations can effectively enhance the structural stability and cyclability of cathode materials [13,14]. It has been reported that a Co0.16Zn0.09V2O5·nH2O electrode exhibits an initial capacity of 90 mAh·g−1 at 3.0 A·g−1, and the capacitance retention is 97% after 1000 cycles [15]. In our previous study, Mn ion-doped VO2(B) was used as the electrode of aqueous metal-ion batteries with enhanced electrochemical performance [16]. It is still necessary, however, to verify the mechanism behind the enhancement of stability and cyclability by the metal cation doping process for aqueous metal-ion batteries.
The electrochemical properties of the VO2(B) material are closely associated with its microstructures and internal defects. The effects of vacancy defects in VO2(B) material on the energy bands, phase-transition characteristics and electronic structure were studied by Wang and Cui based on first principles [17,18]. The computational results showed that vacancy defects were related to the local density of electronic states, which could introduce electrons as free carriers and narrow the energy band gap of the VO2(B) cathode material. The positron annihilation technique is highly sensitive and efficient at detecting the evolution of defects in solid materials. It can particularly detect the local electron density and vacancy defects in materials [19].
In this study, considering that the ionic radius of Zn2+ (0.74 Å) is relatively approximate to that of V4+ (0.58 Å), doping with the Zn ion can create defects and improve the rate capability [20]. Thus, the V1−xZnxO2 samples with different Zn ion doping concentrations were fabricated, and the effect of Zn2+-doping on the microstructure and electrochemical performances of V1−xZnxO2 samples was researched. The experimental results indicated that the doped Zn ion could increase the lattice binding volume of V1−xZnxO2 samples and introduce large-sized vacancy defects, thereby improving the electrochemical properties of the cathode material for aqueous metal-ion batteries.

2. Experiment

2.1. Preparation of V1−xZnxO2 Samples

V1−xZnxO2 samples (0.000 ≤ x ≤ 0.030) were prepared by a simple hydrothermal synthesis method. All raw materials were employed immediately without further purification. A moderate vanadium pentoxide (V2O5) with a purity of 99.99% was blended in oxalic acid (H2C2O4) with a purity of 99.0% at a concentration ratio of 1:3 and stirred for 0.5 h in a magnetic agitator. A specific stoichiometric ratio of zinc nitrate nonahydrate (Zn (NO3)2∙9H2O, 99.99%) as the doping source was added into a mixture of deionized water and hydrogen peroxide (H2O2, 30.0%) and stirred in a magnetic stirrer for 0.5 h. The two resulting mixtures were combined and stirred in a magnetic mixer for 1.5 h to form the mixed liquor. Then, the final mixed liquor was removed to a 250 mL Teflon-packaged stainless steel reaction vessel and stored at 180 °C for 50 h. The precipitation was cleaned six times alternately with anhydrous ethanol and deionized water. Finally, the resulting V1−xZnxO2 nanorods were procured by heating at 80 °C for 15 h in a vacuum oven.

2.2. Characterization of V1−xZnxO2 Samples

The determination of the phase structures of the V1−xZnxO2 samples was performed by X-ray powder diffraction (XRD, SmartLab SE, Cu-Kα radiation, λ = 1.5418 Å) within a 2θ scope of 10°–80° with a scan step of 0.02°. The Le Bail method was used to obtain the lattice parameters and mean crystallite size. The microstructure of the samples was analyzed via scanning electron microscopy (FESEM, JSM-7001F) and high-resolution transmission electron microscopy (HRTEM, JSM2100). The positron annihilation lifetime spectra of the samples were studied via fast-fast coincidence lifetime spectrometer, and 22Na was used as the positron radiation source with a strength of approximately 13 μCi. A PATFIT program was employed to analyze the lifetime spectrum of the samples.

2.3. Electrochemical Measurements

The electrochemical performance of V1−xZnxO2 electrodes was investigated using a CHI760E electrochemical workstation in 2 mol/L KOH aqueous solution at room temperature. The V1−xZnxO2 electrodes were produced by mixing the synthesized nanorods, acetylene black and polytetrafluoroethylene with a quantity ratio of 8:1:1 in anhydrous ethanol. Afterwards, the hybrid paste was evenly coated on a kiln-dried nickel foam plate and vacuum-dried at 80 °C for 12 h for the succeeding measurement. The electrochemical performance was evaluated by the three-electrode test method: the prepared V1−xZnxO2 electrode served as the working electrode and the platinum electrode and calomel electrode were employed as counter electrode and reference electrode, respectively [21]. The cyclic voltammetry (CV), galvanostatic charge–discharge (GCD) and electrochemical impedance spectroscopy (EIS) of the V1−xZnxO2 electrodes were measured.

3. Results and Discussion

3.1. Structural Characterizations

The atomic ratio of zinc, vanadium and oxide in the V1−xZnxO2 with different doping concentrations was calculated according to the XPS spectra. As shown in Figure 1, the actual amounts of zinc are slightly less than the calculated amounts in the fabrication process, which means that not all the zinc ion in the precursor solution could be doped into the VO2(B). The XRD patterns of the V1−xZnxO2 (x = 0.000, 0.005, 0.015, 0.030) samples are shown in Figure 2a. All the diffraction peaks of the samples are related to the monoclinic crystal structure of VO2(B) with the space group C2/m (PDF No. 31-1438), and there is no obvious trace of an impurity phase in the samples. The XRD patterns show that the crystalline grains of all V1−xZnxO2 samples germinated preferentially along the (110) orientation. Figure 2b shows an enlarged image of the (110) peaks of the XRD patterns. It is clear that the (110) peaks of the XRD pattern gradually shifted to the lower degree with the increase in Zn ion doping amounts. This may be due to the lattice expansion caused by the replacement of V4+ (0.58 Å) by Zn2+ (0.74 Å) with a larger radius. However, for the (002) and (−401) peaks in the inset of Figure 2b, little shift can be observed. This phenomenon can be ascribed to the lattice distortion of the (110) crystal plane caused by the doping process. In addition, the relative intensities of the peaks show a drop which may be caused by the decrease in crystalline degree with the doping concentration increased. The noticeable decrease in (001) and (002) can be ascribed to the inhibiting effect on the (001) and (002) crystallographic plane of the doping ion into the crystalline structure of VO2(B). The lattice parameters and unit cell volumes of the V1−xZnxO2 samples were obtained by the Le Bail method and are shown in Table 1. It can be noted that the unit cell volumes of the doped V1−xZnxO2 samples became larger than that of the undoped sample; this may be attributed to the increase in the lattice parameters a and c. The calculated lattice parameters show small and random variations, which can be attributed to the fact that the doping concentration was too low to cause a regular change in the lattice parameters. The results show that the Zn2+ doping with a larger radius led to the expansion of the crystalline structure, which may have benefited the electrochemical performance of the V1−xZnxO2 samples.
Figure 3 shows the SEM pictures of the V1xZnxO2 samples. The surface morphologies of all the V1xZnxO2 samples are nanorods. The average grain size of the V1xZnxO2 samples is between 60 nm and 240 nm. For the x = 0.000, 0.005, 0.015 and 0.030 samples, the average diameters are 133.8 nm, 122.6 nm, 120.6 nm and 101.6 nm, respectively. Notably, the average diameter decreases with increasing doping Zn ion concentrations, which means that the doped Zn2+ with larger radius may have an inhibitory effect on the growth of V1xZnxO2 nanorods.
The crystal structure of the prepared samples can be further analyzed by TEM. Figure 4 shows the TEM images of V1xZnxO2 samples with the doping concentrations of 0.000, 0.005 and 0.015. It can be seen from Figure 4a–c that there is no distinct change in the structure and morphology of the samples after Zn ion doping. Compared with the undoped sample, the doped samples maintain the morphology of nanorods with diameters of about 80–200 nm. This is consistent with the SEM images of Figure 3. In addition, the lattice fringes can be seen clearly in Figure 4d–f, and the high-resolution TEM images display the distances of 0.248 nm, 0.506 nm and 0.312 nm, corresponding to the interplanar distance of the (−402), (−201) and (−202) planes, respectively, which indicate good crystallinity for all samples. The results are consistent with the XRD analysis results in Figure 2 and Table 1.
The lifetime parameters of the positron annihilation of the V1xZnxO2 samples were analyzed by the PATFIT program. The different types of the defects in the samples had different local electron densities, and so corresponded to different positron annihilation lifetimes τ. The relative strength I of the lifetime constituent represented the corresponding concentration of the defects in the materials.
The short lifetime component τ1 is commonly attributed to the positron annihilation in surface states. The long lifetime component τ2 is related to the positron annihilation of defect regions and denotes the size of defects in samples. The intensity I2 reflects the concentration of defects in the samples [22]. As shown in Figure 5, when the doping concentration x of the Zn ion increases from 0.000 to 0.005, the value of τ2 increases linearly. However, when the Zn ion concentration rises from 0.005 to 0.030, the value of τ2 reduces gradually. This phenomenon is caused by the competition between the low-valence Zn ion doping effect and the large-radius Zn ion doping effect on the positron lifetime [23]. To be specific, the larger-radius ion doping will increase the defect size, while the low-valence ion doping will lead to a decrease in the vacancy defect size. When the doping amount of Zn ions is between 0.000 and 0.005, the larger-radius effect is the main factor in determining the defect size. When the doping amount of Zn ions is between 0.005 and 0.015, the low-valence effect is the main factor in determining the defect size. In summary, under the joint modulation of the larger-radius effect and the low-valence effect, the size of defects in the V1xZnxO2 samples increase initially and decrease afterwards with the increase in the Zn doping amount.
The positron bulk lifetime (τb) reflects the local electron density in the crystal lattice and can be computed according to the formula derived from the two-state trapping model [24]:
τb = 1/(I1/τ1 + I2/τ2)
Table 2 shows the positron lifetime values τ1, τ2, τb, τm, I1 and I2 of the V1xZnxO2 samples. The types of defects are judged by the ratio of τ2/τb. When the ratio of τ2/τb is in the range of 1.1–1.3, 1.3–1.4, and >1.5, it represents single vacancy, double vacancy and vacancy clusters, respectively [22]. The computed τ2/τb values are 1.56, 1.71, 1.61 and 1.55 for x = 0.000, 0.005, 0.015 and 0.030 samples, respectively, indicating that vacancy clusters are present in all V1xZnxO2 samples. The mean lifetime τm can be computed according to the formula:
τm = τ1 × I1 + τ2 × I2
This reflects the process of positron annihilation in free states and trap states, and gives detailed information on the electron density and defect distribution inside the material [25]. The τm values of the experimental samples are shown in Table 2. Compared with the undoped VO2(B) sample, the τm of those doped samples becomes significantly larger. This reflects the fact that Zn doping leads to a decrease in the electron density of the vacancy defects and annihilation sites in the lattices, which can be considered as a variation of the chemical environment around the annihilation states.

3.2. Electrochemical Performance

The potential applications of V1−xZnxO2 samples in electrochemical performance were further investigated by the cyclic voltammetry (CV), electrochemical impedance spectroscopy (EIS) and galvanostatic charge–discharge (GCD) in a three-electrode system. The CV curves of V1−xZnxO2 at a scanning rate of 100 mV /s are shown in Figure 6a. The two kinds of peaks, reduction peaks and oxidation peaks, can be seen clearly for all the samples in the CV curves, indicating the electrode process battery behavior. The good reversibility and structural stability of the V1−xZnxO2 electrode can be proved by the symmetrical shape of the CV curves, which indicates the stable (de)intercalation of metal ion in the electrode materials [26,27,28].
As shown in Figure 6b, according to the GCD curves at a current density of 0.1 A∙g−1, the slopes in the voltage range of 0.52 V–0.44 V and 0.32 V–0.20 V are larger, while the slopes in the range of 0.44 V–0.32 V are clearly gentler. When x increases from 0 to 0.015, the discharge time gradually increases, and as x increases to 0.030, the charge and discharge time decreases, and the x = 0.015 sample possesses the best charge–discharge performance.
According to the galvanostatic discharge curves, the specific capacitance (C) of electrodes can be calculated by the following equation:
C = I m · t 3.6
where I (A) represents the load current; m (g) shows the quality of active materials; ∆t (s) represents the discharge time [29,30]. As can be seen from Figure 6c, with an increase in Zn ion-doping concentrations, the calculated specific capacitance of the prepared samples firstly increases and then decreases under the same current density. Compared with the VO2(B) electrode, the specific capacitance performance of the Zn-doped samples is significantly enhanced. In particular, the sample of x = 0.015 possesses the highest specific capacity.
As the most important indicator to evaluate electrochemical performance, the integrated electrochemical impedance spectroscopy (EIS) is used to judge the electrode conductivity, charge transfer performance and diffusion property [31]. The Nyquist plot for V1−xZnxO2 electrodes within the frequency range of 10−2–105 Hz is shown in Figure 6d. The inset (1) of Figure 6d shows the simulated equivalent circuit, where Rs denotes the electrolyte resistance, Rct indicates the resistance of charge transfer between the electrolyte and electrode, Cdl shows the double-layer capacitor representing the contact electrode surface and W shows the Warburg impedance [32,33]. In the low frequency range, there is no significant difference in the slopes of the impedance curves for all samples. As shown in the inset (2) of Figure 6d, the intersection points of the impedance curve with the x axis signify the ohmic resistance between the electrolyte and the surface of the electrode materials. It is obvious that the doped samples possess lower resistance. When the doping amount x is 0.015, the resistance is the lowest, indicating the highest electrochemical activity, which is in accord with the CV and GCD results.
According to the above electrochemical results, it can be seen that the V1−xZnxO2 (x = 0.015) sample has the best electrochemical performance. In order to further explore the enhancement mechanism of the electrochemical activity behind the doping process, we conducted more detailed electrochemical measurements on the V1−xZnxO2 (x = 0.015) sample. The CV curves at different scan rates are illustrated in Figure 7a, from which we can see that the potential window increases with the increase in the scan rate, but still maintains a similar shape. These phenomena can be ascribed to the excellent amplification performance of the V1−xZnxO2 (x = 0.015) electrode. Figure 7b demonstrates the CV curves at 100 mV/s for 1000 cycles. The peak shape of the CV curves has good repeatability after 1000 cycles. This indicates that V1−xZnxO2 (x = 0.015) still maintains good cycling stability. Figure 7c shows the GCD curves of V1−xZnxO2 (x = 0.015) electrode at different current densities. The GCD curves exhibit the behavioral features of a supercapacitor, which are primarily attributed to the quasi-reversible redox reactions or electrochemical desorption/adsorption on the electrode contact surface in the electrolyte. In addition, it can be seen in Figure 7d that the capacitance retention is 88% after 1000 cycles. This means the V1−xZnxO2 (x = 0.015) electrode possesses excellent cycling durability, which may be caused by its crystalline structural stability. The electrochemical performance of the V1−xZnxO2 (x = 0.015) electrode indicates its potential application as electrodes for energy storage devices.

4. Conclusions

In summary, the influence of Zn ion-doping on the crystalline structure, morphologies and defect formation has carefully explored. The potential applications on electrochemical performance have also been studied. The doping process can cause the lattice expansion in the unit cell volume. SEM and TEM images indicate that all the prepared samples are well-crystallized nanorods. The results of positron lifetime spectroscopy confirm that Zn ion-doping can introduce large vacancy clusters. Electrochemical measurements demonstrate that the electrochemical performance of V1−xZnxO2 samples can be enhanced by doped Zn ions. This work can provide an effective protocol for the fabrication of advanced electrode materials with optimized performance.

Author Contributions

Conceptualization, D.L., X.Z. and X.W.; Formal analysis, Q.Z. and X.C.; Funding acquisition, D.L., H.D., X.Z., J.C., G.G. and C.S.; Project administration, D.L. and X.W.; Supervision, D.L., H.D., X.Z., J.C., G.G., C.S. and X.W.; Writing—original draft, Q.Z.; Writing—review and editing, D.L. and X.W. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the National Natural Science Foundation of China (Grant Nos. 12005194, 11804311, 11775192, 11405148), the Key Research & Development and promotion projects in Henan Province (Grant Nos. 212102210132, 212102210477, 212300410092, 222102230101) and the postgraduate education reform and quality improvement projects in Henan Province (Grant Nos. HNYJS2021AL027, DWJZW202131zn).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bella, F.; Griffini, G.; Correa, J.P.; Saracco, G.; Gratzel, M.; Hagfeldt, A.; Turri, S.; Gerbaldi, C. Improving efficiency and stability of perovskite solar cells with photocurable fluoropolymers. Science 2016, 354, 203–206. [Google Scholar] [CrossRef] [PubMed]
  2. Yang, Z.G.; Zhang, J.L.; Kintner-Meyer, M.C.W.; Lu, X.C.; Choi, D.W.; Lemmon, J.P.; Liu, J. Electrochemical energy storage for green grid. Chem. Rev. 2011, 111, 3577–3613. [Google Scholar] [CrossRef] [PubMed]
  3. Huie, M.M.; Bock, D.C.; Takeuchi, E.S.; Marschilok, A.C.; Takeuchi, K.J. Cathode materials for magnesium and magnesium-ion based batteries. Coord. Chem. Rev. 2015, 287, 15–27. [Google Scholar]
  4. Liu, Q.; Hu, Z.; Zhang, Y.; Xing, G.; Tang, Y.; Chou, S. Designing Advanced Vanadium-Based Materials to Achieve Electrochemically Active Multielectron Reactions in Sodium/Potassium-Ion Batteries. Adv. Energy Mater. 2020, 10, 2002244. [Google Scholar]
  5. Liu, J.; Xu, C.; Chen, Z.; Ni, S.; Shen, Z. Progress in aqueous rechargeable batteries. Green Energy Environ. 2018, 3, 20–41. [Google Scholar] [CrossRef]
  6. Chen, X.; Wang, L.; Li, H.; Cheng, F.; Chen, J. Porous V2O5 nanofibers as cathode materials for rechargeable aqueous zinc-ion batteries. J. Energy Chem. 2019, 38, 20–25. [Google Scholar] [CrossRef] [Green Version]
  7. Zhang, Y.; Chen, A.; Sun, J. Promise and challenge of vanadium-based cathodes for aqueous Zinc-ion batteries. J. Energy Chem. 2021, 54, 655–667. [Google Scholar] [CrossRef]
  8. Wan, F.; Niu, Z. Design Strategies for Vanadium-based Aqueous Zinc-Ion Batteries. Angew. Chem. Int. Ed. 2019, 58, 16358–16367. [Google Scholar] [CrossRef]
  9. Wei, Q.; Liu, J.; Feng, W.; Sheng, J.; Tian, X.; He, L.; An, Q.; Mai, L. Hydrated vanadium pentoxide with superior sodium storage capacity. J. Mater. Chem. A 2015, 3, 8070. [Google Scholar] [CrossRef]
  10. Wang, W.; Jiang, B.; Hu, L.; Lin, Z.; Hou, J.; Jiao, S. Single crystalline VO2 nanosheets: A cathode material for sodi-um-ion batteries with high rate cycling performance. J. Power Source 2014, 250, 181. [Google Scholar] [CrossRef]
  11. Liu, Y.; Wu, X. Review of vanadium-based electrode materials for rechargeable aqueous zinc ion batteries. J. Energy Chem. 2021, 56, 223–237. [Google Scholar] [CrossRef]
  12. Zhao, Q.; Jiao, L.; Peng, W.; Gao, H.; Yang, J.; Wang, Q.; Du, H.; Li, L.; Qi, Z.; Si, Y.; et al. Facile synthesis of VO2(B)/carbon nanobelts with high capacity and good cyclability. J. Power Source 2012, 199, 350–354. [Google Scholar] [CrossRef]
  13. Wang, F.; Liu, Y.; Liu, C.Y. Hydrothermal synthesis of carbon/vanadium dioxide core–shell microspheres with good cycling performance in both organic and aqueous electrolytes. Electrochim. Acta 2010, 55, 2662–2666. [Google Scholar] [CrossRef]
  14. Liu, G.; Du, Y.; Liu, W.; Wen, L. Study on the action mechanism of doping transitional elements in spinel LiNi0.5Mn1.5O4. Electrochim. Acta 2016, 209, 308–314. [Google Scholar] [CrossRef]
  15. Huang, H.; Tian, T.; Pan, L.; Chen, X.; Tervoort, E.; Shih, C.J.; Niederberger, M. Layered metal vanadates with different interlayer cations for high-rate Na-ion storage. J. Mater. Chem. A 2019, 7, 16109. [Google Scholar] [CrossRef] [Green Version]
  16. Liu, D.; Zhang, Q.; Ding, S.; Yan, F.; Dai, H.; Li, T.; Xue, R.; Chen, J.; Gong, G.; Shang, C.; et al. Microdefects evolution and electrochemical performance modulation of Mn doped VO2(B) nanorods. J. Alloys Compd. 2022, 911, 164975. [Google Scholar] [CrossRef]
  17. Fang, D.L.; Li, J.C.; Liu, X.; Huang, P.F.; Xu, T.R.; Qian, M.C.; Zheng, C.H. Synthesis of a Co-Ni doped LiMn2O4 spinel cathode material for high-power Li-ion batteries by a sol-gel mediated solid-state route. J. Alloys Compd. 2015, 640, 82–89. [Google Scholar] [CrossRef]
  18. Akkila, T.; Mansur Basha, I. Study of structural, morphological and optical properties of nano-structured zinc doped V2O5 thin films. J. Pure Appl. Sci. Technol. 2018, 8, 18–33. [Google Scholar]
  19. Cui, Y.Y.; Liu, B.; Chen, L.L.; Luo, H.J.; Gao, Y.F. Formation energies of intrinsic point defects in monoclinic VO2 studied by first-principles calculations. AIP Adv. 2016, 6, 105301–105309. [Google Scholar] [CrossRef] [Green Version]
  20. Wang, X.; Wang, Z.; Zhang, G.; Jiang, J. Insight into electronic and structural reorganizations for defect-induced VO2 metal-insulator transition. J. Phys. Chem. Lett. 2017, 8, 3129–3132. [Google Scholar] [CrossRef]
  21. Kumar, N.S.; Chang, J.H.; Ho, M.-S.; Balraj, B.; Chandrasekar, S.; Mohanbabu, B.; Gowtham, M.; Guo, D.; Mohanraj, K. Impact of Zn2+ Doping on the Structural, Morphological and Photodiode Properties of V2O5 Nanorods. J. Inorg. Organomet. Polym. Mater. 2020, 31, 1066–1078. [Google Scholar] [CrossRef]
  22. Dai, H.; Ye, F.; Chen, Z.; Li, T.; Liu, D. The effect of ion doping at different sites on the structure, defects and multiferroic properties of BiFeO3 ceramics. J. Alloys Compd. 2018, 734, 60–65. [Google Scholar] [CrossRef]
  23. Qi, X.; Dho, J.; Tomov, R.; Blamire, M.G.; MacManus-Driscoll, J.L. Greatly reduced leakage current and conduction mechanism in aliovalent-ion-doped BiFeO3. Appl. Phys. Lett. 2005, 86, 062903. [Google Scholar] [CrossRef]
  24. Puska, M.J.; Nieminen, R.M. Theory of positrons in solids and on solid surfaces. Rev. Mod. Phys. 1994, 66, 841. [Google Scholar] [CrossRef] [Green Version]
  25. Wang, M.; Dai, H.; Li, T.; Chen, J.; Yan, F.; Xue, R.; Xing, X.; Chen, D.; Ping, T.; He, J. The evolution of structure and properties in GdMn(1−x)TixO3 ceramics. J. Mater. Sci. Mater. Electron. 2021, 32, 27348–27361. [Google Scholar] [CrossRef]
  26. Wang, H.; Wang, M.; Tang, Y. A novel zinc-ion hybrid supercapacitor for long-life and low-cost energy storage applications. Energy Storage Mater. 2018, 13, 1–7. [Google Scholar] [CrossRef]
  27. Wu, X.; Tao, Y.; Dong, L.; Wang, Z.; Hu, Z. Preparation of VO2 nanowires and their electric characterization. Mater. Res. Bull. 2005, 40, 315–321. [Google Scholar] [CrossRef]
  28. Li, N.; Huang, W.; Shi, Q.; Zhang, Y.; Song, L. A CTAB-assisted hydrothermal synthesis of VO2 (B) nanostructures for lithium-ion battery application. Ceram. Int. 2013, 39, 6199–6206. [Google Scholar] [CrossRef]
  29. Wang, X.; Xi, B.; Ma, X.; Feng, Z.; Jia, Y.; Feng, J.; Qian, Y.; Xiong, S. Boosting Zinc-Ion Storage Capability by Effectively Suppressing Vanadium Dissolution Based on Robust Layered Barium Vanadate. Nano Lett. 2020, 20, 2899–2906. [Google Scholar] [CrossRef]
  30. Ma, Y.; Zhu, X.; Wang, B.; Liu, S.; Meng, T.; Chen, H.; Peng, B.; Deng, Z. Sacrificial template synthesis of hierarchical nickel hydroxidenitrate hollow colloidal particles for electrochemical energy storage. Chem. Eng. Sci. 2020, 217, 115548. [Google Scholar] [CrossRef]
  31. Yang, H.; Cheng, Z.; Wu, P.; Wei, Y.; Jiang, J.; Xu, Q. Deep eutectic solvents regulation synthesis of multi-metal oxalate for electrocatalytic oxygen evolution reaction and supercapacitor applications. Electrochim. Acta 2022, 427, 140879. [Google Scholar] [CrossRef]
  32. Mahadi, N.B.; Park, J.S.; Park, J.H.; Chung, K.Y.; Yi, S.Y.; Sun, Y.K.; Myung, S.T. Vanadium dioxide-reduced graphene oxide composite as cathode materials for rechargeable Li and Na batteries. J. Power Source 2016, 326, 522–532. [Google Scholar] [CrossRef]
  33. Liu, Q.; Tan, G.Q.; Wang, P.; Abeyweera, S.C.; Zhang, D.T.; Rong, Y.C.; Wu, Y.A.; Lu, J.; Sun, C.J.; Ren, Y.; et al. Revealing mechanism responsible for structural reversibility of single-crystal VO2 nanorods upon lithiation/delithiation. Nano Energy 2017, 36, 197–205. [Google Scholar] [CrossRef]
Figure 1. (a) XPS spectra of V1−xZnxO2 samples; (b) XPS spectra of Zn 2p3/2 and Zn 2p1/2; the inset of (b) is the atomic ratio of oxide, vanadium and zinc.
Figure 1. (a) XPS spectra of V1−xZnxO2 samples; (b) XPS spectra of Zn 2p3/2 and Zn 2p1/2; the inset of (b) is the atomic ratio of oxide, vanadium and zinc.
Batteries 08 00172 g001
Figure 2. (a) XRD patterns of V1−xZnxO2 samples; (b) the enlarged view of (110) peaks of the XRD patterns; the inset of (b) is the enlarged view of (002) and (−401) peaks.
Figure 2. (a) XRD patterns of V1−xZnxO2 samples; (b) the enlarged view of (110) peaks of the XRD patterns; the inset of (b) is the enlarged view of (002) and (−401) peaks.
Batteries 08 00172 g002
Figure 3. SEM images of V1xZnxO2 samples: (a) x = 0.000; (b) x = 0.005; (c) x = 0.015; (d) x = 0.030.
Figure 3. SEM images of V1xZnxO2 samples: (a) x = 0.000; (b) x = 0.005; (c) x = 0.015; (d) x = 0.030.
Batteries 08 00172 g003
Figure 4. TEM and HRTEM images of V1−xZnxO2 samples: (a) and (d) x = 0.000; (b) and (e) x = 0.005; (c) and (f) x = 0.015; the inset of (d) is an enlarged view of the lattice fringes.
Figure 4. TEM and HRTEM images of V1−xZnxO2 samples: (a) and (d) x = 0.000; (b) and (e) x = 0.005; (c) and (f) x = 0.015; the inset of (d) is an enlarged view of the lattice fringes.
Batteries 08 00172 g004
Figure 5. The positron lifetime parameters τ1, τ2 and I2 of V1−xZnxO2 samples.
Figure 5. The positron lifetime parameters τ1, τ2 and I2 of V1−xZnxO2 samples.
Batteries 08 00172 g005
Figure 6. Electrochemical measurements of V1−xZnxO2 samples: (a) CV curves at different doping concentrations; (b) Potential vs. time for charge-discharge profiles of all samples at 0.1 A/g; (c) Specific capacitance of samples with different doping concentrations; (d) Nyquist plots curves.
Figure 6. Electrochemical measurements of V1−xZnxO2 samples: (a) CV curves at different doping concentrations; (b) Potential vs. time for charge-discharge profiles of all samples at 0.1 A/g; (c) Specific capacitance of samples with different doping concentrations; (d) Nyquist plots curves.
Batteries 08 00172 g006
Figure 7. Electrochemical measurement of V1xZnxO2 samples with x = 0.015: (a) CV curves at different scan rates; (b) the CV curves after 1000 cycles at a scan rate of 100 mV/s; (c) galvanostatic discharge profiles at different current density; (d) cycling performance at 1 A·g−1.
Figure 7. Electrochemical measurement of V1xZnxO2 samples with x = 0.015: (a) CV curves at different scan rates; (b) the CV curves after 1000 cycles at a scan rate of 100 mV/s; (c) galvanostatic discharge profiles at different current density; (d) cycling performance at 1 A·g−1.
Batteries 08 00172 g007
Table 1. The calculated lattice constants and cell volumes of the V1−xZnxO2 samples.
Table 1. The calculated lattice constants and cell volumes of the V1−xZnxO2 samples.
Doping ConcentrationLattice Parametersβ
xa (Å)b (Å)c (Å)V (Å3)
0.00012.07073.69586.4270274.23106.966
0.00512.08013.69436.4411275.02106.909
0.01512.07593.69816.4319274.81106.906
0.03012.08423.69256.4552275.57106.912
Table 2. The positron lifetime values τ1, I1, τ2, I2, τb and τm of V1−xZnxO2 samples.
Table 2. The positron lifetime values τ1, I1, τ2, I2, τb and τm of V1−xZnxO2 samples.
SamplePositron Lifetime/psIntensity/%Bulk Lifetime/psMean Lifetime/ps
xτ1τ2I1I2τbτm
0.000279.4468.482.817717.1823300.2141311.8745
0.005319.9574.289.153210.8468336.0428347.4834
0.015322.1543.189.111510.8885337.0332346.1636
0.030327.7537.286.011813.9882346.6081357.0053
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, D.; Zhang, Q.; Chen, X.; Dai, H.; Zhai, X.; Chen, J.; Gong, G.; Shang, C.; Wang, X. Microstructure Modulation of Zn Doped VO2(B) Nanorods with Improved Electrochemical Properties towards High Performance Aqueous Batteries. Batteries 2022, 8, 172. https://doi.org/10.3390/batteries8100172

AMA Style

Liu D, Zhang Q, Chen X, Dai H, Zhai X, Chen J, Gong G, Shang C, Wang X. Microstructure Modulation of Zn Doped VO2(B) Nanorods with Improved Electrochemical Properties towards High Performance Aqueous Batteries. Batteries. 2022; 8(10):172. https://doi.org/10.3390/batteries8100172

Chicago/Turabian Style

Liu, Dewei, Qijie Zhang, Xiaohong Chen, Haiyang Dai, Xuezhen Zhai, Jing Chen, Gaoshang Gong, Cui Shang, and Xuzhe Wang. 2022. "Microstructure Modulation of Zn Doped VO2(B) Nanorods with Improved Electrochemical Properties towards High Performance Aqueous Batteries" Batteries 8, no. 10: 172. https://doi.org/10.3390/batteries8100172

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop