Next Article in Journal
Investigating the Anticancer Potential of Zinc and Magnesium Alloys: From Base Materials to Nanocoated Titanium Implants
Previous Article in Journal
Development of Carbon Black Coating on TPU Elastic Powder for Selective Laser Sintering
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Electronic Properties and Mechanical Stability of Multi-Ion-Co-Intercalated Bilayered V2O5

Institute of Modern Physics, Shaanxi Key Laboratory for Theoretical Physics Frontiers, Northwest University, Xi’an 710127, China
*
Author to whom correspondence should be addressed.
Materials 2024, 17(13), 3364; https://doi.org/10.3390/ma17133364
Submission received: 22 May 2024 / Revised: 27 June 2024 / Accepted: 3 July 2024 / Published: 8 July 2024

Abstract

:
Incorporating metal cations into V2O5 has been proven to be an effective method for solving the poor long-term cycling performance of vanadium-based oxides as electrodes for mono- or multivalent aqueous rechargeable batteries. This is due to the existence of a bilayer structure with a large interlayer space in the V2O5 electrode and to the fact that the intercalated ions act as pillars to support the layered structure and facilitate the diffusion of charged carriers. However, a fundamental understanding of the mechanical stability of multi-ion-co-intercalated bilayered V2O5 is still lacking. In this paper, a variety of pillared vanadium pentoxides with two types of co-intercalated ions were studied. The root-mean-square deviation of the V-O bonds and the elastic constants calculated by density functional theory were used as references to evaluate the stability of the intercalated compounds. The d-band center and electronic band structures are also discussed. Our theoretical results show that the structural characteristics and stability of the system are quite strongly influenced by the intercalating strategy.

1. Introduction

Lithium-ion batteries (LIBs) are extensively used in the market because of their superior high-energy density [1,2]. However, their future growth has been limited by several key issues, such as resource issues and security. In recent years, aqueous rechargeable batteries (ARBs) with alkaline cations (Li+, Na+, and K+) [3,4,5,6,7] and multivalent cations (Zn2+, Mg2+, and Al3+) [8,9,10] have attracted increasing interest and are considered powerful alternatives to LIBs. There are numerous reports on cathode materials, such as vanadium oxides [11,12,13,14,15], manganese-based materials [16], and Prussian blue [17,18]. Among them, vanadium oxides have drawn attention owing to their flexible multivalent state, remarkable specific capacity, and abundant reserves in nature.
V2O5 is one of the few electrode materials that are able to accommodate both monovalent and multivalent cations within a layered structure. The layers are connected by van der Waals forces and allow for the (de)intercalation of various ions. Aqueous electrolytes and hydrated vanadium oxide (VOH) electrodes are used in aqueous V2O5 batteries. The presence of H2O modifies the α-V2O5 structure into a bilayer structure, which has a larger interlayer space, thus facilitating the diffusion of metal ions [14]. The bilayer structure is typical in xerogels, nanosheets, and nanotubes. It was further discovered that the insertion of metal ions can improve the electrode performance by flexibly adjusting the layer spacing and reinforce the stability of VOH by supporting the V–O layer, with the ions acting like pillars. Investigators explored the regulation of the VOH structure after various metal ions were inserted between the VOH layers, such as Na+, Ca2+, Mg2+, and Al3+ [12,19,20,21]. Recent studies have found that the organic molecule polyaniline (PANI) can also be inserted between the layers of VOH, resulting in an increase in the layer spacing to 14.08 Å and maintaining the stability of the layered structure [22]. In recent years, the pre-intercalation strategy has also been increasingly applied in zinc batteries. For example, by intercalating polyaniline and water in V2O5 nanowires, significant capacity gains and improved cycle stability were achieved [23]. Zhang et al. constructed a vanadium oxide hydrate with an tunable P-band center for zinc batteries through a pre-intercalation strategy [24]. Li et al. proposed a 3D flower-like NH4V4O10 micro-morphology with preinserted Mg2+ ions as a cathode material for AZIBs, exhibiting a high specific capacity, a satisfactory rate performance, and an excellent life cycle [25]. Although these methods can effectively improve the electrochemical performance of vanadium oxide, metal ions are small in volume and have limited ability to expand layers. Alternatively, co-intercalating, i.e., inserting two or more kinds of extrinsic atoms simultaneously, is considered an effective method to improve the interlayer space and performance of the electrode. Until now, many metal ion pre-intercalation strategies have focused on V2O5-based cathodes [11,12,13,14,15]. In both single-ion and multiple-ion pre-intercalation, the same fundamental structural unit of the V2O5 bilayer supported by two different types of ions is used and plays an essential role in determining the structural stability and electrochemical performance of the electrodes. However, there has been no specialized theoretical study focusing on this structural pattern. The different roles of alkali and transition metal cations in improving the stability and the electrochemical performance of vanadium-based materials are still unclear. The advantages and disadvantages of the pillars have not been identified. It is meaningful to fully investigate the effect of co-intercalating pairs composed of common alkali metals and transition metal ions.
This paper focuses on co-pre-inserted combinations of monovalent and divalent/trivalent cations in bilayered V2O5 as cathode materials for multivalent ion batteries. We mainly discuss the structural stability and electronic properties of bilayered V2O5 after metal ion intercalation. Our results may open new opportunities for the development of high-performance vanadium oxide-based cathodes for multivalent-cation batteries.

2. Methodology

Computational Details

All the calculations in this work were performed using density functional theory (DFT) with the Vienna Ab Initio Simulation Package (VASP5.4.4) [26,27]. The projector augmented-wave (PAW) method was employed, which has been widely used for battery materials and has shown good predictive capability [28,29,30]. Generalized gradient approximation (GGA) [31] with Perdew–Burke–Ernzerhof (PBE) parameterization was applied as the exchange-correlation functional. Although Hubbard U correction is usually employed since it accounts for on-site Coulomb interactions for the 3d orbitals of V, it was not included due to the sufficiency of PBE parameterization for voltage calculations and the poor convergence previously observed with the +U correction for diffusion calculations [32]. All calculation parameters were selected after the convergence test of the total energy of the system. The cutoff energy of the plane wave was set to 500 eV. The Brillouin zone of the unit cells was sampled by a 4 × 4 × 4 mesh according to the Monkhorst–Pack scheme. The energy and the force convergence criterion were chosen to be 10−7 eV and 10−3 eV Å−1, respectively.
In order to investigate the thermodynamic stability of the structure after intercalation, the formation energy was calculated. The formation energy is defined as follows:
E f o r m a t i o n = E A B V 8 O 20 E V 8 O 20 E A E B
where E A B V 8 O 20 is the total energy of the co-intercalated system, E V 8 O 20 is the energy of the bilayer V2O5 crystal, and E A and E B are the energy per atom of metal ions A/B, respectively, obtained from bulk metallic calculations. With this definition, a negative E f o r m a t i o n means that the intercalating process is exothermic, and a large absolute value means increased stability.
In addition to the thermodynamic stability, the mechanical stability of the intercalated crystals was studied. To judge the stability of the intercalated crystals, elastic stability conditions were adopted [33,34]. For the crystal axes XX-1, YY-2, ZZ-3, XY(YX)-4, YZ(ZY)-5, and XZ(ZX)-6, Voigt representation was adopted, and the axis arrangement is shown in Figure 1. Bilayer V2O5 belongs to the monoclinic crystal system. When two cations are introduced, its crystal system remains within the monoclinic category. There are 13 independent elastic constants associated with it, and the elastic constant matrix is diagonal. The stability criteria of the monoclinic crystal system are as follows [35,36]:
C 11 > 0 ,   C 22 > 0 ,   C 33 > 0 ,   C 44 > 0 ,   C 55 > 0 ,   C 66 > 0
( C 22 + C 33 2 C 23 ) > 0 ,   [ C 11 + C 22 + C 33 + 2 ( C 12 + C 13 + C 23 ) ] > 0
( C 33 C 55 C 35 2 ) > 0 ,   ( C 44 C 66 C 46 2 ) > 0
[ C 22 ( C 33 C 55 C 35 2 ) + 2 C 23 C 25 C 35 C 23 2 C 55 C 25 2 C 33 ] > 0
g = C 11 C 22 C 33 C 11 C 23 2 C 22 C 13 2 C 33 C 12 2 + 2 C 12 C 13 C 23
[ 2 C 15 C 25 ( C 33 C 12 C 13 C 23 ) + C 15 C 35 ( C 22 C 13 C 12 C 23 ) + C 25 C 35 ( C 11 C 23 C 12 C 13 ) C 15 2 ( C 22 C 33 C 23 2 ) + C 25 2 ( C 11 C 33 C 13 2 ) + C 35 2 ( C 11 C 22 C 12 2 ) + C 55 g ] > 0

3. Results and Discussion

3.1. Structure Modification after Co-Intercalation

The structural instability of the common orthorhombic α-V2O5 (space group Pmmn) phase has emerged as a constraint for multivalent-cation battery design. The interplanar spacing between the [VO5−VO5] polyhedral chains in α-V2O5 may not be sufficient to reversibly accommodate large-ionic-radius cations within its crystal volume. Meanwhile, the presence of large 2D channels for ion transport and the very thin solid network of V2O5·nH2O-derived materials stimulate further investigation regarding their potential application as electrode materials in multivalent cation batteries. X-ray diffraction (XRD) measurements revealed that V2O5·nH2O xerogels are composed of bilayers of V2O5 (monoclinic, space Group C2/m) separated by water layers [37]. The bilayer V2O5 comprises two layers, each with the V2O5 stoichiometry facing each other, which confers distinct electrochemical behavior to the V2O5 gel. Special focus is being put on incorporating metal cations into bilayer V2O5 to improve the capacity and cyclic stability and maintain the bilayer structure of the V2O5 frame [38,39,40,41,42].
Compared to the single-cation intercalation strategy, a co-intercalating strategy with two types of cations could be more attractive. Considering economics and practicality, we chose alkali metals (Li, Na, K), alkaline earth metals (Mg, Ca) and transition metals in the first four cycles (Sc, Ti, V, Cr, Mn, Fe, Co, Ni, Cu, Zn). A 1 × 2 × 1 supercell of bilayer V2O5 was used, and we fixed the intercalation positions of intercalation atoms A and B at two symmetric positions, (0.125, 0.5, 0.5) and (0.875, 0.5, 0.5), as suggested for ν-V2O5 and δ-V2O5, which are analogous to bilayer V2O5 [43]. The cell parameters and atomic positions were further optimized by means of DFT calculations, as shown in Figure 1. One A and one B ion were intercalated into the bilayer V2O5, giving a stoichiometric formula of ABV8O20. Our theoretical model is reasonable considering the pre-intercalated metal ions in bilayer V2O5 that have previously been investigated, such as NH4V4O10 [44], Mg0.34V2O5 [12], and Na0.33V2O5 [19].
Water molecules are essential in the electrochemical process of aqueous rechargeable batteries [45]. Structural water, a common pre-intercalated molecule, can expand the layer spacing and effectively shield the charge interaction of metal ions. In published works, water molecules have been shown to have a positive effect on the stability of the electrode. This is not the main factor in the instability of the electrode. The hydrated cation and proton dynamics are complicated and are beyond the scope of this manuscript.
The optimized unit cell parameters for bilayer V2O5 are a = 11.69 Å, b = 3.61Å, c = 9.85 Å, α = 90°, β = 96.47°, and γ = 90°, which are in reasonable agreement with the experimental values reported in Ref. [43]. The mean deviation is about 1.5%. The optimized lattice parameters of pure bilayered V2O5 obtained using different methods are shown in Table 1 [46]. With the introduction of metal ions, the layer spacing of V2O5 decreases due to the strong ionic bonds with the vanadium oxide bronze. As the structures both before and after intercalation exhibit clear layering, the c-axis lattice constant was chosen in this paper as an indicator of layer spacing. As illustrated in Figure 2, the incorporation of Mg (ionic radius 72 pm) [47], Ca (100 pm), or Sc (75 pm) ions may decrease the layer spacing due to their strong correlation with the V-O framework, while Al (53 pm) and K (151 pm) may enlarge the layer spacing, accompanied with significant structural changes.
In order to gain further insights into the intercalated structure and estimate the deviation of the co-intercalated structure from the pristine one, the root-mean-square deviation (drms) of the 32 V-O bonds involved in our model was calculated. The maximum deviation of the V-O bond lengths (dmax) is also shown in Figure 2. The apex of the square pyramid is a vanadyl bond (V-O) that is shorter (1.5–1.6 Å) than the other V-O bonds (1.8–2.1 Å). The root-mean-square deviation of the 32 V-O bonds serves as a reference value to evaluate the structural stability after ion intercalation. We took 0.1 Å as the lower limit of drms, which suggests potential damage to the layered structure. This standard is not mandatory.
For the Li-doped cases, as shown in Figure 2a, the combinations of (Li, Na), (Li, K), (Li, Ca), and (Li, Ni) exhibited slight deviations from pristine bilayered V2O5, which indicates stability with pillars supporting the bilayer structure. For the Na-doped cases, the (Na, Li), (Na, K), and (Na, Ni) pairs exhibited a small deviation, which can be judged from the drms and dmax values. Additionally, co-intercalations of (K, Ni), (K, Cu), (Ca, K), (Ca, Co), (Ca, Ni), and (Ca, Cu) showed a minimal influence on the bilayer structure.
Although the ionic size of Al3+ (0.53Å) [47] is smaller than that of Li+ (0.76 Å), the high charge density of trivalent Al3+ cations can lead to strong interactions between the Al3+ cations and the framework of the main structure, potentially resulting in structural changes or slow kinetics during the process of Al3+ intercalation. The incorporation of aluminum ions dramatically changes the bilayer structure in each case. Also, the drms value is relatively large, more than 0.1 Å. With the involvement of Al3+, the bilayer V2O5 suffers from structural phase transformation. Our theoretical results indicate a similar trend for cases involving Sc and Ti ions.
Bilayer V2O5 is widely used as a cathode for aqueous zinc-ion batteries (ZIBs). Mai’s group reported Na0.33V2O5 (NVO) with enhanced performance (253.7 mA h g−1 @ 200 mAg−1) [19]. Husam N. Alshareef’s group also demonstrated that a better ZIB performance could be achieved on Ca0.25V2O5·nH2O (CVO) (340 mA h g−1 @ 50 mAg−1) [20] and Mg0.34V2O5·nH2O (340 mA h g−1 @ 100 mA g−1) [12]. Liang’s group further developed several kinds of potassium vanadate nanobelts (K2V8O21, K0.25V2O5, K2V6O16·1.57H2O, and KV3O8) [49] for ZIBs. The root-mean-square deviation (drmx) values for (Na, Zn), (Ca, Zn), (Mg, Zn), and (K, Zn) are 0.041, 0.048, 0.047, and 0.047 Å, which are smaller than the 0.1 Å standard for stability analysis. Shuhua Wang’s group reported a reversible Zn//(Na, Mn)V8O20·nH2O system (377 mA h g−1@ 100 mA g−1), achieved by introducing manganese (Mn) ions into NaV8O20 [8]. In Wang’s work, the c-axis lattice constant of (Na0.43Mn0.53)V8O20·nH2O was 10.62 Å, while that of (Na, Mn)V8O20 obtained by our calculations was 9.36 Å. We attribute this difference to the fact that our system does not consider the unique role of water in expanding the layer spacing. A drms value of 0.043 Å was obtained for the (Na, Mn) system, indicating that this co-intercalating strategy can maintain the bilayer structure and facilitate the diffusion of Zn ions. Our stability analysis based on the root-mean-square deviation of V-O bonds is consistent with these reported metal-ion-pre-intercalated V2O5 experiments.

3.2. Stability and Mechanical Properties after Co-Intercalation

To provide further insight into the stability of the bilayer V2O5 after cation intercalation, we also calculated the formation energies, shown in Figure 3. All the co-intercalated systems have negative formation energy, indicating that they are thermodynamically stable and can easily be synthesized by sol–gel chemistry or hydrothermal methods.
To further estimate the mechanical stability of V2O5 bilayers with various guests, the elastic constants of these materials were calculated. The diagonal elastic constants Cii (C11, C22, C33) are dominant and are listed in Table 2. Based on these elastic constants, there are 62 stable pairs and 13 unstable pairs among the 75 co-intercalated pairs considered. The unstable co-intercalated systems that did not meet the stability criterion, as outlined in the methodology section, were marked as “*”.
The combinations (Li, Al), (Na, Al), (Al, Sc), (Al, Cu), and (Al, Zn)-V8O20 involving Al3+ are unstable, which is consistent with the aforementioned V-O bond deviation analysis. This may be related to the serious structural deformation of (Al, B)-V8O20, which leads to damage to the bilayer structure. We noticed a recent work by Q. Pang et al. that presented aluminum-pre-intercalated V2O5 as a high-performance cathode material for aqueous zinc-ion batteries [21]. Their powder X-ray diffraction (XRD) study showed an orthorhombic V2O5 structure, while no monoclinic bilayer structure was detected. In a recent work by P. De et al., two-dimensional V2O5 nanosheets were reported as a cathode material for realizing low-cost aqueous aluminum-ion batteries. The formation of layered orthorhombic V2O5 with the Pmmn space group was also confirmed by XRD diffraction [50].
The mechanical properties of an electrode are also an essential consideration in the design of mono- and multivalent cation batteries. Assuming that electrode particles are typically polycrystalline and can be modeled as isotropic elastic materials, the averaged bulk (B), shear (G), and Young’s (E) moduli and Poisson’s ratio (ν) can be obtained from the Cij values based on Reuss’s lower bound, Voigt’s upper bound, and Hill’s homogenization schemes [51]. These values were calculated with the help of the software VASPKIT (1.5.0) [52]. The calculated results are shown in Table 2, and the Voigt approximation was still adopted. The bulk modulus, Young’s modulus, and shear modulus of the pristine bilayer V2O5 were 64.09, 69.43, and 26.31 GPa. The data in Table 2 suggest that the stiffness of the material is enforced by co-intercalation.
According to the criterion of fracture behavior, materials with a low Poisson’s ratio are brittle materials [53]. A Poisson’s ratio of ν < 0.26 indicates that the material is brittle, while ν > 0.26 indicates that the material is ductile [54]. Ductile materials are expected to be used as electrode materials for flexible batteries. The elastic constant C33 relates to interlayer interactions, which also have a strong influence on the flexibility of the electrode. The smallest C33 was observed for (Mg, Ti) co-intercalation, while the largest value was from the (Al, Cr) case, indicating a denser structure.
The electrode materials can be stiffened by Ca intercalation, as judged from the elastic constants. More attention should be paid to K ions, since the formation energy is negative, the structural deviation after the intercalation is relatively small, the interlayer spacing is overall large compared with the other cases, and the d-band center indicates lesser H2 generation with all cases involving K. K ions seem to be good candidates as pillars for bilayer V2O5.
Due to the rich structural chemistry of vanadium oxide frameworks, our work gives ideas of interest for research on V2O5-based electrode materials for battery applications. However, some points should be further investigated, such as the interactions between the intercalating sites and the role of structural water in assisting cation insertion. Continued efforts are underway to solve these problems. A validated “cocktail method” for vanadium-based electrode material design may appear after further studies.

3.3. Electronic Structure

Density functional theory (DFT) calculations were carried out to study the electronic band structure of vanadium oxide before and after guest intercalation. The main characteristic of the band structure of V2O5-based material is the existence of split bands separated from higher conduction bands, which are weakly dispersing due to V dxy orbitals overlapping with O 2p. Due to the intercalation of alkali ions, the split-off bands of most of the intercalation configurations are partially occupied, leading to charge ordering and interesting 1D magnetic properties. The intercalated V2O5 electrodes show good electrical conductivity overall. We also noticed that the (Mg, Ca), (Mg, Sc), (Mg, Cr), (Ca, Cr), (Ca, Mn), and (Ca, Zn)-V8O20 systems are obvious semiconductors with a small band gap, which indicates poor electrical conductance, as shown in Table 2.
Water splitting in aqueous environments should be considered when designing materials for advanced ARBs [55]. Water decomposition will not only cause electrolyte consumption, resulting in gas expansion of the battery and safety problems, but also consume electrons and reduce the coulomb efficiency of the battery [45]. This process involves two half-reactions of the oxygen evolution reaction (OER) and hydrogen evolution reaction (HER), which exhibit different reaction mechanisms in materials with different d-band centers. The d-band center is an effective indicator of hydrogen solution and diffusion [56]. A shift upward (downward) of the d-band center enhances (weakens) hydrogen solution ability. The d-band centers of the (Al, Mn), (Al, Ti)-V8O20 pairs are positioned above the Fermi energy, which warrants further attention. The (Li, Na), (Al, Cr), and (Al, V) pairs tend to boost hydrogen generation, while the (Li, K), (Na, Sc), and (Ca, Sc) pairs suppress such generation, as illustrated in Figure 3.

4. Conclusions

Two-dimensional bilayer V2O5, which offers the advantage of a more open framework for ion intercalation, is an attractive cathode material for aqueous rechargeable batteries. Multi-ion co-pre-intercalation is an effective method used to modify electrodes, stabilize layered structures, and improve electrochemical performance. Co-intercalation cases involving two types of ions were studied systematically. This theoretical study mainly focused on the changes after ion intercalation in terms of the bond length, electronic structure, stability, and mechanical properties. The calculation results showed that the (Li, Al), (Na, Al), (Na, Sc), (Na, Co), (Na, Ni), (Mg, Ti), (Mg, Fe), (Al, Sc), (Al, Cu). (Al, Zn), (K, Ca), (Ca, Co), and (Ca, Ni) pairs are not stable. In contrast, the (Li, Na), (Na, K), (Li, K), and (K, Cu) pairs presented a small deviation from the pristine layered structure. As judged from the root-mean-square deviation of the V-O bonds, the K ion can enlarge the interlayer space and maintain the layer structure. Incorporating Al, Ti, Sc, and V harms the bilayer structure. Calculations of the d-band center revealed that the (Li, K), (Na, Sc), and (Ca, Sc) pairs can suppress water splitting on the surface of the electrode. Our work provides more co-intercalation schemes and insights into vanadium oxide-based electrode materials.

Author Contributions

C.M.: validation, writing—original draft preparation; B.Z.: methodology, writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Double First-class University Construction Project of Northwest University, National Natural Science Foundation of China under Grant Nos. 11947301 and 12047502.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The raw data supporting the conclusions of this article will be made available by the authors on request.

Conflicts of Interest

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.

References

  1. Zhang, F.; Qi, L. Recent Progress in Self-Supported Metal Oxide Nanoarray Electrodes for Advanced Lithium-Ion Batteries. Adv. Sci. 2016, 3, 1600049. [Google Scholar] [CrossRef] [PubMed]
  2. Zhang, X.-D.; Shi, J.-L.; Liang, J.-Y.; Yin, Y.-X.; Guo, Y.-G.; Wan, L.-J. Structurally modulated Li-rich cathode materials through cooperative cation doping and anion hybridization. Sci. China Chem. 2017, 60, 1554–1560. [Google Scholar] [CrossRef]
  3. Zhang, Z.; Wang, H.; Ji, S.; Pollet, B.G.; Wang, R. V2O5-SiO2 hybrid as anode material for aqueous rechargeable lithium batteries. Ionics 2016, 22, 1593–1601. [Google Scholar] [CrossRef]
  4. Champness, N.R. The future of metal-organic frameworks. Dalton Trans. 2011, 40, 10311–10315. [Google Scholar] [CrossRef] [PubMed]
  5. Qu, Q.T.; Liu, L.L.; Wu, Y.P.; Holze, R. Electrochemical behavior of V2O5·0.6H2O nanoribbons in neutral aqueous electrolyte solution. Electrochim. Acta 2013, 96, 8–12. [Google Scholar] [CrossRef]
  6. Xue, L.; Li, Y.; Lin, W.; Chen, F.; Chen, G.; Chen, D. Electrochemical properties and facile preparation of hollow porous V2O5 microspheres for lithium-ion batteries. J. Colloid Interface Sci. 2023, 638, 231–241. [Google Scholar] [CrossRef] [PubMed]
  7. Reddy, I.N.; Akkinepally, B.; Manjunath, V.; Neelima, G.; Reddy, M.V.; Shim, J. SnO2 Quantum Dots Distributed along V2O5 Nanobelts for Utilization as a High-Capacity Storage Hybrid Material in Li-Ion Batteries. Molecules 2021, 26, 7262. [Google Scholar] [CrossRef] [PubMed]
  8. Du, M.; Liu, C.; Zhang, F.; Dong, W.; Zhang, X.; Sang, Y.; Wang, J.-J.; Guo, Y.-G.; Liu, H.; Wang, S. Tunable Layered (Na, Mn)V8O20·nH2O Cathode Material for High-Performance Aqueous Zinc Ion Batteries. Adv. Sci. 2020, 7, 2000083. [Google Scholar] [CrossRef] [PubMed]
  9. Wang, F.; Fan, X.; Gao, T.; Sun, W.; Ma, Z.; Yang, C.; Han, F.; Xu, K.; Wang, C. High-Voltage Aqueous Magnesium Ion Batteries. ACS Cent. Sci. 2017, 3, 1121–1128. [Google Scholar] [CrossRef]
  10. Wang, H.; Bai, Y.; Chen, S.; Luo, X.; Wu, C.; Wu, F.; Lu, J.; Amine, K. Binder-free V2O5 cathode for greener rechargeable aluminum battery. ACS Appl. Mater. Interfaces 2015, 7, 80–84. [Google Scholar] [CrossRef]
  11. Liu, C.; Neale, Z.; Zheng, J.; Jia, X.; Huang, J.; Yan, M.; Tian, M.; Wang, M.; Yang, J.; Cao, G. Expanded hydrated vanadate for high-performance aqueous zinc-ion batteries. Energy Environ. Sci. 2019, 12, 2273–2285. [Google Scholar] [CrossRef]
  12. Ming, F.; Liang, H.; Lei, Y.; Kandambeth, S.; Eddaoudi, M.; Alshareef, H.N. Layered MgxV2O5·nH2O as Cathode Material for High-Performance Aqueous Zinc Ion Batteries. ACS Energy Lett. 2018, 3, 2602–2609. [Google Scholar] [CrossRef]
  13. Sa, N.; Kinnibrugh, T.L.; Wang, H.; Sai Gautam, G.; Chapman, K.W.; Vaughey, J.T.; Key, B.; Fister, T.T.; Freeland, J.W.; Proffit, D.L.; et al. Structural Evolution of Reversible Mg Insertion into a Bilayer Structure of V2O5·nH2O Xerogel Material. Chem. Mater. 2016, 28, 2962–2969. [Google Scholar] [CrossRef]
  14. Yan, M.; He, P.; Chen, Y.; Wang, S.; Wei, Q.; Zhao, K.; Xu, X.; An, Q.; Shuang, Y.; Shao, Y.; et al. Water-Lubricated Intercalation in V2O5·nH2O for High-Capacity and High-Rate Aqueous Rechargeable Zinc Batteries. Adv. Mater. 2018, 30, 1703725. [Google Scholar] [CrossRef] [PubMed]
  15. Yang, Y.; Tang, Y.; Fang, G.; Shan, L.; Guo, J.; Zhang, W.; Wang, C.; Wang, L.; Zhou, J.; Liang, S. Li+ intercalated V2O5·nH2O with enlarged layer spacing and fast ion diffusion as an aqueous zinc-ion battery cathode. Energy Environ. Sci. 2018, 11, 3157–3162. [Google Scholar] [CrossRef]
  16. Sun, W.; Wang, F.; Hou, S.; Yang, C.; Fan, X.; Ma, Z.; Gao, T.; Han, F.; Hu, R.; Zhu, M.; et al. Zn/MnO2 Battery Chemistry With H+ and Zn2+ Coinsertion. J. Am. Chem. Soc. 2017, 139, 9775–9778. [Google Scholar] [CrossRef] [PubMed]
  17. Yang, Q.; Mo, F.; Liu, Z.; Ma, L.; Li, X.; Fang, D.; Chen, S.; Zhang, S.; Zhi, C. Activating C-Coordinated Iron of Iron Hexacyanoferrate for Zn Hybrid-Ion Batteries with 10 000-Cycle Lifespan and Superior Rate Capability. Adv. Mater. 2019, 31, 1901521. [Google Scholar] [CrossRef] [PubMed]
  18. Deng, L.; Yang, Z.; Tan, L.; Zeng, L.; Zhu, Y.; Guo, L. Investigation of the Prussian Blue Analog Co3[Co(CN)6]2 as an Anode Material for Nonaqueous Potassium-Ion Batteries. Adv. Mater. 2018, 30, e1802510. [Google Scholar] [CrossRef]
  19. He, P.; Zhang, G.; Liao, X.; Yan, M.; Xu, X.; An, Q.; Liu, J.; Mai, L. Sodium Ion Stabilized Vanadium Oxide Nanowire Cathode for High-Performance Zinc-Ion Batteries. Adv. Energy Mater. 2018, 8, 1702463. [Google Scholar] [CrossRef]
  20. Xia, C.; Guo, J.; Li, P.; Zhang, X.; Alshareef, H.N. Highly Stable Aqueous Zinc-Ion Storage Using a Layered Calcium Vanadium Oxide Bronze Cathode. Angew. Chem. Int. Ed. Engl. 2018, 57, 3943–3948. [Google Scholar] [CrossRef]
  21. Pang, Q.; He, W.; Yu, X.; Yang, S.; Zhao, H.; Fu, Y.; Xing, M.; Tian, Y.; Luo, X.; Wei, Y. Aluminium pre-intercalated orthorhombic V2O5 as high-performance cathode material for aqueous zinc-ion batteries. Appl. Surf. Sci. 2021, 538, 148043. [Google Scholar] [CrossRef]
  22. Kuchena, S.F.; Wang, Y. V2O5 intercalated with polyaniline for improved kinetics in aqueous ammonium-ion batteries. Electrochim. Acta 2022, 425, 140751. [Google Scholar] [CrossRef]
  23. Zhu, Y.; Liu, X.; Hu, X.; Wang, T.; Parkin, I.P.; Wang, M.; Boruah, B.D. Polyaniline and water pre-intercalated V2O5 cathodes for high-performance planar zinc-ion micro-batteries. Chem. Eng. J. 2024, 487, 150384. [Google Scholar] [CrossRef]
  24. Feng, Z.; Zhang, Y.; Gao, Z.; Hu, D.; Jiang, H.; Hu, T.; Meng, C.; Zhang, Y. Construction interlayer structure of hydrated vanadium oxides with tunable P-band center of oxygen towards enhanced aqueous Zn-ion batteries. Adv. Powder Mater. 2024, 3, 100167. [Google Scholar] [CrossRef]
  25. Wang, X.; Wang, Y.; Naveed, A.; Li, G.; Zhang, H.; Zhou, Y.; Dou, A.; Su, M.; Liu, Y.; Guo, R.; et al. Magnesium Ion Doping and Micro-Structural Engineering Assist NH4V4O10 as a High-Performance Aqueous Zinc Ion Battery Cathode. Adv. Funct. Mater. 2023, 33, 2306205. [Google Scholar] [CrossRef]
  26. Kresse, G.; Furthmüller, J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 1996, 6, 15–50. [Google Scholar] [CrossRef]
  27. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169–11186. [Google Scholar] [CrossRef] [PubMed]
  28. Bréger, J.; Meng, Y.S.; Hinuma, Y.; Kumar, S.; Kang, K.; Shao-Horn, Y.; Ceder, G.; Grey, C.P. Effect of High Voltage on the Structure and Electrochemistry of LiNi0.5Mn0.5O2: A Joint Experimental and Theoretical Study. Chem. Mater. 2006, 18, 4768–4781. [Google Scholar] [CrossRef]
  29. Zhou, F.; Kang, K.; Maxisch, T.; Ceder, G.; Morgan, D. The electronic structure and band gap of LiFePO4 and LiMnPO4. Solid State Commun. 2004, 132, 181–186. [Google Scholar] [CrossRef]
  30. Kang, K.; Morgan, D.; Ceder, G. First principles study of Li diffusion in I-Li2NiO2 structure. Phys. Rev. B 2009, 79, 014305. [Google Scholar] [CrossRef]
  31. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [PubMed]
  32. Sandagiripathira, K.; Moghaddasi, M.A.; Shepard, R.; Smeu, M. Investigating the role of structural water on the electrochemical properties of α-V2O5 through density functional theory. Phys. Chem. Chem. Phys. 2022, 24, 24271–24280. [Google Scholar] [CrossRef] [PubMed]
  33. Dobson, P.J.J.P.B. Physical Properties of Crystals—Their Representation by Tensors and Matrices. Phys. Bull. 1985, 36, 506. [Google Scholar] [CrossRef]
  34. Mouhat, F.; Coudert, F.-X. Necessary and sufficient elastic stability conditions in various crystal systems. Phys. Rev. B 2014, 90, 224104. [Google Scholar] [CrossRef]
  35. Gao, X.P.; Jiang, Y.H.; Liu, Y.Z.; Zhou, R.; Feng, J. Stability and elastic properties of NbxCy compounds. Chin. Phys. B 2014, 23, 097704. [Google Scholar] [CrossRef]
  36. Zhou, B.; Shi, H.; Cao, R.; Zhang, X.; Jiang, Z. Theoretical study on the initial stage of a magnesium battery based on a V2O5 cathode. Phys. Chem. Chem. Phys. 2014, 16, 18578–18585. [Google Scholar] [CrossRef] [PubMed]
  37. Kristoffersen, H.H.; Metiu, H. Structure of V2O5·nH2O Xerogels. J. Phys. Chem. C 2016, 120, 3986–3992. [Google Scholar] [CrossRef]
  38. Liu, Z.; Sun, H.; Qin, L.; Cao, X.; Zhou, J.; Pan, A.; Fang, G.; Liang, S. Interlayer Doping in Layered Vanadium Oxides for Low-cost Energy Storage: Sodium-ion Batteries and Aqueous Zinc-ion Batteries. ChemNanoMat 2020, 6, 1553–1566. [Google Scholar] [CrossRef]
  39. Guo, S.; Fang, G.; Liang, S.; Chen, M.; Wu, X.; Zhou, J. Structural perspective on revealing energy storage behaviors of silver vanadate cathodes in aqueous zinc-ion batteries. Acta Mater. 2019, 180, 51–59. [Google Scholar] [CrossRef]
  40. Shan, L.; Yang, Y.; Zhang, W.; Chen, H.; Fang, G.; Zhou, J.; Liang, S. Observation of combination displacement/intercalation reaction in aqueous zinc-ion battery. Energy Storage Mater. 2019, 18, 10–14. [Google Scholar] [CrossRef]
  41. Yu, X.; Hu, F.; Cui, F.; Zhao, J.; Guan, C.; Zhu, K. The displacement reaction mechanism of the CuV2O6 nanowire cathode for rechargeable aqueous zinc ion batteries. Dalton Trans. 2020, 49, 1048–1055. [Google Scholar] [CrossRef] [PubMed]
  42. Zheng, J.; Liu, C.; Tian, M.; Jia, X.; Jahrman, E.P.; Seidler, G.T.; Zhang, S.; Liu, Y.; Zhang, Y.; Meng, C.; et al. Fast and reversible zinc ion intercalation in Al-ion modified hydrated vanadate. Nano Energy 2020, 70, 104519. [Google Scholar] [CrossRef]
  43. Galy, J. Vanadium pentoxide and vanadium oxide bronzes—Structural chemistry of single (S) and double (D) layer MxV2O5 phases. J. Solid State Chem. 1992, 100, 229–245. [Google Scholar] [CrossRef]
  44. Li, Q.; Rui, X.; Chen, D.; Feng, Y.; Xiao, N.; Gan, L.; Zhang, Q.; Yu, Y.; Huang, S. A High-Capacity Ammonium Vanadate Cathode for Zinc-Ion Battery. Nano-Micro Lett. 2020, 12, 67. [Google Scholar] [CrossRef] [PubMed]
  45. Li, M.; Li, Z.; Wang, X.; Meng, J.; Liu, X.; Wu, B.; Han, C.; Mai, L. Comprehensive understanding of the roles of water molecules in aqueous Zn-ion batteries: From electrolytes to electrode materials. Energy Environ. Sci. 2021, 14, 3796–3839. [Google Scholar] [CrossRef]
  46. Wu, T.; Zhu, K.; Qin, C.; Huang, K. Unraveling the role of structural water in bilayer V2O5 during Zn2+-intercalation: Insights from DFT calculations. J. Mater. Chem. A 2019, 7, 5612–5620. [Google Scholar] [CrossRef]
  47. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr. Sect. A 1976, 32, 751–767. [Google Scholar] [CrossRef]
  48. McColl, K.; Corà, F. Phase stability of intercalated V2O5 battery cathodes elucidated through the Goldschmidt tolerance factor. Phys. Chem. Chem. Phys. 2019, 21, 7732–7744. [Google Scholar] [CrossRef]
  49. Tang, B.; Fang, G.; Zhou, J.; Wang, L.; Lei, Y.; Wang, C.; Lin, T.; Tang, Y.; Liang, S. Potassium vanadates with stable structure and fast ion diffusion channel as cathode for rechargeable aqueous zinc-ion batteries. Nano Energy 2018, 51, 579–587. [Google Scholar] [CrossRef]
  50. De, P.; Halder, J.; Priya, S.; Srivastava, A.K.; Chandra, A. Two-Dimensional V2O5 Nanosheets as an Advanced Cathode Material for Realizing Low-Cost Aqueous Aluminum-Ion Batteries. ACS Appl. Energy Mater. 2023, 6, 753–762. [Google Scholar] [CrossRef]
  51. Liu, Y.; Jiang, Y.; Feng, J.; Zhou, R. Elasticity, electronic properties and hardness of MoC investigated by first principles calculations. Phys. B Condens. Matter 2013, 419, 45–50. [Google Scholar] [CrossRef]
  52. Wang, V.; Xu, N.; Liu, J.-C.; Tang, G.; Geng, W.-T. VASPKIT: A user-friendly interface facilitating high-throughput computing and analysis using VASP code. Comput. Phys. Commun. 2021, 267, 108033. [Google Scholar] [CrossRef]
  53. Ding Ying-Chun, X.B. Electronic Structure, Mechanical Properties and Intrinsic Hardness of a New Superhard Material BeP2N4. Acta Phys. Chim. Sin. 2011, 27, 1621–1632. [Google Scholar]
  54. Vaitheeswaran, G.; Kanchana, V.; Svane, A.; Delin, A. Elastic properties of MgCNi3—a superconducting perovskite. J. Phys. Condens. Matter 2007, 19, 326214. [Google Scholar] [CrossRef]
  55. Ma, L.; Li, N.; Long, C.; Dong, B.; Fang, D.; Liu, Z.; Zhao, Y.; Li, X.; Fan, J.; Chen, S.; et al. Achieving Both High Voltage and High Capacity in Aqueous Zinc-Ion Battery for Record High Energy Density. Adv. Funct. Mater. 2019, 29, 1906142. [Google Scholar] [CrossRef]
  56. Zhu, Q.; Huang, W.; Huang, C.; Gao, L.; Su, Y.; Qiao, L. The d band center as an indicator for the hydrogen solution and diffusion behaviors in transition metals. Int. J. Hydrog. Energy 2022, 47, 38445–38454. [Google Scholar] [CrossRef]
Figure 1. The structure of AB-V8O20. The orange sphere represents the intercalated atom A, the blue sphere represents the intercalated atom B, and gray and red spheres represent vanadium and oxygen atoms, respectively.
Figure 1. The structure of AB-V8O20. The orange sphere represents the intercalated atom A, the blue sphere represents the intercalated atom B, and gray and red spheres represent vanadium and oxygen atoms, respectively.
Materials 17 03364 g001
Figure 2. Structural changes under different intercalation combinations. (af) Structural changes in Li-, Na-, Mg-, Al-, K-, and Ca-dominated combinations. The bar chart shows the change in bond length, and the broken line chart shows the change in layer spacing.
Figure 2. Structural changes under different intercalation combinations. (af) Structural changes in Li-, Na-, Mg-, Al-, K-, and Ca-dominated combinations. The bar chart shows the change in bond length, and the broken line chart shows the change in layer spacing.
Materials 17 03364 g002
Figure 3. Formation energy and d-band center of different intercalation combinations.
Figure 3. Formation energy and d-band center of different intercalation combinations.
Materials 17 03364 g003
Table 1. Calculated lattice parameters of pure V2O5 obtained using different methods.
Table 1. Calculated lattice parameters of pure V2O5 obtained using different methods.
D-V2O5a (Å)b (Å)c (Å)
PBE (this work)11.693.619.85
DFT-PBE-D3 [48]11.363.569.37
PBE + U [46]11.583.658.59
Table 2. Elastic constants and elastic modulus for different intercalation situations. The Voigt notation replaces XX→1, YY→2, ZZ→3, ZY(YZ)→4, XZ(ZX)→5, and XY(YX)→6. All values are in units of GPa.
Table 2. Elastic constants and elastic modulus for different intercalation situations. The Voigt notation replaces XX→1, YY→2, ZZ→3, ZY(YZ)→4, XZ(ZX)→5, and XY(YX)→6. All values are in units of GPa.
C11C22C33B (GPa)E (GPa)G (GPa)v
Li, Na175.26 220.71 41.20 86.30 84.67 31.68 0.34 metal
Li, Mg176.67 223.58 39.78 92.33 75.35 27.62 0.36 metal
Li, Al *137.08 229.34 44.59 72.64 63.91 23.61 0.35 metal
Li, K167.36 214.76 34.34 81.77 86.38 32.62 0.32 metal
Li, Ca191.37 242.23 43.84 99.24 96.45 36.04 0.34 metal
LI, Sc172.38 225.90 89.33 99.16 90.72 33.66 0.35 metal
Li, Ti107.21 235.99 96.20 89.31 82.64 30.71 0.35 metal
Li, V117.16 240.22 100.75 91.08 88.82 33.20 0.34 metal
Li, Cr181.39 238.00 73.39 102.71 97.23 36.22 0.34 metal
Li, Mn181.03 242.48 64.85 100.64 90.84 33.65 0.35 metal
Li, Fe174.04 230.18 48.86 95.77 76.12 27.83 0.37 metal
Li, Co175.50 227.10 27.49 84.06 82.85 31.01 0.34 metal
Li, Ni180.49 225.17 35.08 92.08 88.26 32.93 0.34 metal
Li, Cu174.10 229.84 39.47 89.79 84.08 31.28 0.34 metal
Li, Zn183.26 227.57 45.33 92.43 86.73 32.28 0.34 metal
Na, Mg192.11 229.62 52.22 95.88 94.63 35.43 0.34 metal
Na, Al *134.19 213.54 27.69 67.29 70.94 26.78 0.32 metal
Na, K163.82 210.07 35.39 80.41 84.65 31.96 0.32 metal
Na, Ca176.04 236.56 45.93 93.15 92.19 34.53 0.34 metal
Na, Sc *168.74 227.96 42.56 86.82 76.12 28.11 0.35 metal
Na, Ti132.90 236.19 64.50 85.31 83.13 31.07 0.34 metal
Na, V139.34 224.61 81.76 81.30 91.10 34.69 0.31 metal
Na, Cr188.66 231.72 67.35 99.89 98.81 37.01 0.34 metal
Na, Mn192.04 228.80 51.41 95.43 94.42 35.36 0.34 metal
Na, Fe184.22 220.59 41.70 95.18 86.56 32.10 0.35 metal
Na, Co *182.06 235.02 27.00 78.03 81.14 30.58 0.33 metal
Na, Ni *171.86 229.76 62.91 79.50 90.02 34.32 0.31 metal
Na, Cu173.65 223.93 47.30 89.93 80.87 29.95 0.35 metal
Na, Zn191.81 225.46 48.41 95.09 93.25 34.89 0.34 metal
Mg, Al181.36 257.38 62.95 99.18 103.06 38.84 0.33 metal
Mg, K179.63 225.05 39.73 90.08 100.97 38.45 0.31 metal
Mg, Ca207.43 248.33 68.19 108.05 111.40 41.94 0.33 0.16
Mg, Sc190.59 202.86 101.40 93.06 95.58 35.97 0.33 0.19
Mg, Ti *178.62 224.56 16.78 84.77 69.08 25.32 0.36 metal
Mg, V215.85 255.24 89.67 116.66 100.84 37.20 0.36 metal
Mg, Cr216.94 252.68 63.09 113.40 97.13 35.78 0.36 0.15
Mg, Mn175.28 252.18 59.21 99.06 90.10 33.41 0.35 metal
Mg, Fe *192.59 240.38 39.68 101.32 70.01 25.28 0.38 metal
Mg, Co155.68 216.09 36.41 72.94 77.57 29.32 0.32 metal
Mg, Ni178.05 243.94 63.53 101.67 88.47 32.65 0.35 metal
Mg, Cu176.35 228.65 47.06 95.33 81.21 29.90 0.36 metal
Mg, Zn138.83 181.10 25.19 63.24 68.06 25.77 0.32 metal
Al, K144.54 226.89 30.74 74.99 94.67 36.71 0.29 metal
Al, Ca190.85 248.94 110.96 113.22 112.94 42.34 0.33 metal
Al, Sc *177.47 220.54 86.65 95.23 77.35 28.34 0.36 metal
Al, Ti190.12 278.23 75.31 102.18122.2547.000.30metal
Al, V174.82 265.06 63.13 99.00 112.96 43.12 0.31 metal
Al, Cr203.13 305.86 144.41 133.42 155.83 59.69 0.31 metal
Al, Mn140.36 219.45 43.54 71.9378.7929.900.32metal
Al, Fe169.46 249.93 56.82 98.69 97.10 36.34 0.34 metal
Al, Co182.68 261.93 66.51 101.06 108.13 40.91 0.32 metal
Al, Ni154.21 253.78 43.35 83.72 85.90 32.32 0.33 metal
Al, Cu *143.62 219.17 31.41 70.33 53.84 19.61 0.37 metal
Al, Zn *118.89 204.58 33.62 66.12 66.29 24.87 0.33 metal
K, Ca *156.52 219.61 54.43 87.04 62.61 22.68 0.38 metal
K, Sc154.01 223.96 46.59 81.08 85.07 32.10 0.33 metal
K, Ti116.71 221.52 89.44 77.60 94.77 36.55 0.30 metal
K, V155.35 231.92 101.99 92.75 104.50 39.82 0.31 metal
K, Cr186.73 227.42 54.76 96.52 106.68 40.54 0.32 metal
K, Mn187.15 223.47 45.57 92.77 101.68 38.59 0.32 metal
K, Fe166.40 209.33 34.01 81.24 87.45 33.11 0.32 metal
K, Co164.14 220.29 40.81 83.72 91.57 34.75 0.32 metal
K, Ni167.36 217.93 37.09 86.53 87.23 32.74 0.33 metal
K, Cu176.80 225.94 38.47 89.41 84.04 31.28 0.34 metal
K, Zn170.14 218.06 36.94 83.59 89.45 33.84 0.32 metal
Ca, Sc190.92 249.69 66.62 106.57 108.02 40.58 0.33 metal
Ca, Ti191.04 240.23 132.97 112.17 109.79 41.06 0.34 metal
Ca, V194.87 244.95 130.29 114.00 116.08 43.63 0.33 metal
Ca, Cr201.92 251.90 94.50 113.01 117.09 44.11 0.33 0.17 †
Ca, Mn205.68 248.31 66.70 107.45 110.49 41.58 0.33 0.13
Ca, Fe205.88 252.55 75.81 110.53 111.69 41.94 0.33 metal
Ca, Co *178.11 251.16 68.70 103.27 105.75 39.78 0.33 metal
Ca, Ni *183.65 244.71 83.99 105.95 90.26 33.23 0.36 metal
Ca, Cu177.89 246.19 68.00 99.96 94.01 34.99 0.34 metal
Ca, Zn208.13 247.74 70.06 108.41 110.38 41.49 0.33 0.19
* represents the unstable pairs; † represents the direct band gap.
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

Ma, C.; Zhou, B. Electronic Properties and Mechanical Stability of Multi-Ion-Co-Intercalated Bilayered V2O5. Materials 2024, 17, 3364. https://doi.org/10.3390/ma17133364

AMA Style

Ma C, Zhou B. Electronic Properties and Mechanical Stability of Multi-Ion-Co-Intercalated Bilayered V2O5. Materials. 2024; 17(13):3364. https://doi.org/10.3390/ma17133364

Chicago/Turabian Style

Ma, Chunhui, and Bo Zhou. 2024. "Electronic Properties and Mechanical Stability of Multi-Ion-Co-Intercalated Bilayered V2O5" Materials 17, no. 13: 3364. https://doi.org/10.3390/ma17133364

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