Next Article in Journal
Influence of Breathing and Swelling on the Jelly-Roll Case Gap of Cylindrical Lithium-Ion Battery Cells
Next Article in Special Issue
An Overview of Challenges and Strategies for Stabilizing Zinc Anodes in Aqueous Rechargeable Zn-Ion Batteries
Previous Article in Journal
Building Bridges: Unifying Design and Development Aspects for Advancing Non-Aqueous Redox-Flow Batteries
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Defect Chemistry in Zn3V4(PO4)6

by
Navaratnarajah Kuganathan
Department of Materials, Faculty of Engineering, Imperial College London, London SW7 2AZ, UK
Batteries 2023, 9(1), 5; https://doi.org/10.3390/batteries9010005
Submission received: 19 October 2022 / Revised: 6 December 2022 / Accepted: 19 December 2022 / Published: 22 December 2022
(This article belongs to the Special Issue Zinc-Ion Batteries: Issues and Opportunities)

Abstract

:
Zinc-ion batteries have attracted great interest for their low cost, safety, and high energy density. Recently, Zn3V4(PO4)6 has been reported to be a promising cathode material for zinc-ion batteries. The defect chemistry, diffusion of Zn-ions, and solution of dopants are examined by advanced simulation techniques. The simulation results show that the most favorable intrinsic defect is the Zn-V anti-site. A zig-zag pattern of long-range Zn2+ diffusion is observed and the activation energy of 1.88 eV indicates that the ionic conductivity of this material is low. The most promising isovalent dopants on the Zn site are Ca2+ and Fe2+. Although the solution of Ga3+, Sc3+, In3+, Y3+, Gd3+, and La3+ on the V site is exoergic, the most promising is In3+. Different reaction routes for the formation of Zn3V4(PO4)6 are considered and the most thermodynamically favorable reaction consists of binary oxides (ZnO, V2O3, and P2O5) as reactants.

1. Introduction

The battery is a promising sustainable energy source for both mobile and stationary applications [1,2,3]. Among various rechargeable batteries, lithium-ion batteries (LIBs) have been utilized for many years in many energy storage devices such as laptops, mobile phones, and e-bikes [4,5,6,7,8,9]. Although LIBs have achieved excellent progress, factors such as limited lithium resources, safety, and cost have led to finding other monovalent ion batteries consisting of Na+ and K+ ions [10,11,12,13] and multivalent ion batteries based on Mg2+, Ca2+, and Zn2+ ions [14,15,16,17,18,19].
Zinc ion batteries (ZIBs) have received extensive attention owing to their high volumetric energy density, low redox potential, low cost, safety, and high abundance [20,21,22,23,24,25,26]. In order to improve energy efficiency, there is a necessity to find new materials that can be used to construct key components of ZIBs (i.e., anodes, cathodes, and electrolytes).
The cathode is an important part of ZIBs hosting Zn2+ ions. The most promising cathode material that has attracted many researchers is MnO2 due to its crystal structure and different oxidation states of Mn [27,28]. Other cathode materials studied so far include vanadium oxides [29], Prussian blue analogs [30], spinel-structured oxides [31], and organic materials [32].
Phosphate-based materials have attracted great attention for constructing cathodes due to their structural rigidity provided via strong [PO4]3− units and high operating voltages [33]. NASICON-type phosphate materials (e.g., Li3V2(PO4)3 and Na3V2(PO4)3) [34,35], fluorophosphates (e.g., Na3V2(PO4)2F3) [36], olivine-structure-based phosphates (e.g., LiFePO4) [37], and layered phosphates (e.g., VOPO4) [38] are expected to be candidate cathode materials for ZIBs. Recently, a vanadium-based phosphate material, Zn3V4(PO4)6 [39], was reported to be a candidate cathode material for ZIBs owing to its high stability and specific capacity. The performance of this material depends also on the diffusivity of Zn2+ ions in this material, and such diffusion relies on the defects. Although defects and diffusion properties of Zn3V4(PO4)6 are experimentally challenging, computational modeling techniques can provide valuable information about the most thermodynamically stable defect processes, diffusion pathways and their activation energies, and solution of dopants. In a recent classical simulation study of spinel-ZnCo2O4 [40], it was reported that the Co-Zn anti-site is the most stable defect type and that Zn2+ ion diffusion is three-dimensional with an activation energy of 0.70 eV.
In this study, we used the computational modeling technique based on the classical pair-wise potentials to examine the defect, diffusion, and dopant properties of Zn3V4(PO4)6. Many oxide materials have been modeled using this methodology and the calculated properties agree with those measured in the experiments [41,42,43,44,45]. The electronic structures of the most stable dopants were analyzed using density functional theory (DFT) simulations.

2. Computational Methods

Defect simulations were employed using a classical simulation code GULP (Generalized Utility Lattice Program) [46]. In this simulation technique, interactions between ions in the lattice are described using long-range Coulombic and short-range electron–electron repulsion (Pauli) and attractive dispersion (van der Waals) forces. Short-range interactions were modeled using Buckingham potentials (see Table 1) [47,48]. Both lattice constants and ionic positions were relaxed using the Broyden–Fletcher–Goldfarb–Shanno (BFGS) method [49]. Point defect modeling was carried out by a method proposed by the Mott–Littleton method [50]. Vacancy-mediated diffusion of Zn2+ ions was modeled considering seven interstitial positions between two Zn atomic positions in a local hop. The energy difference was between the maximum energy and the energy of the Zn vacancy. Our simulation techniques assumed that all ions were spherical and fully charged. There was an overestimation of the defect energies but the trend’s consistency was retained.
DFT simulations were employed to examine the electronic structures of the most favorable doped configurations. The plane-wave DFT simulation code VASP (Vienna Ab initio Simulation Package) [51] was used. In all cases, a plane wave basis set consisting of a cut-off of 500 eV, projected augmented wave (PAW) pseudopotentials [52], and a 4 × 4 × 4 Monkhorst–Pack [53] k-point mesh were used. The PBE-GGA (Perdew, Burke, and Ernzerhof-generalized gradient approximation) [54] was used to describe the exchange-correlation energy. Full geometry optimization was performed using the conjugate gradient algorithm [55] with a force tolerance of 0.001 eV/Å. The Bader charge analysis [56] was used to estimate the effective charges on the atoms in the relaxed configurations.

3. Results and Discussion

3.1. Crystal Structure of Zn3V4(PO4)6

Zn3V4(PO4)6 is three-dimensional with a triclinic space group of P 1 ¯ . Its experimental lattice constants were reported to be a = 6.349 Å, b = 7.869 Å, c = 9.324 Å, α = 105.32°, β = 108.66°, and γ = 101.23° [57]. In this complex structure, phosphorous and vanadium form tetrahedral units and octahedral units with nearest-neighbor oxygen atoms (see Figure 1a). These two units share their corners and form networks in the crystal. This complex structure was energy-minimized completely to determine the equilibrium lattice constants. The calculated values were in good agreement with the values reported in the experiment showing the efficacy of potential parameters and pseudopotentials (see Table 2).
The calculated total DOS plot shows that the Zn3V4(PO4)6 exhibits a metallic characteristic (see Figure 1b), which is one of the essential conditions for a cathode material. Figure 1c shows the charge density plot showing the ion–ion interaction in the Zn3V4(PO4)6.
Bader charges calculated on the atoms in the bulk Zn3V4(PO4)6 are reported in Table 3. Based on the formula of the Zn3V4(PO4)6, ionic charges on the Zn, V, P, and O are +2.00, +3.00, +5.00, and −2.00, respectively. Although the Bader charges of Zn, V, and O are not closer to the values of the full ionic charge model, the bulk Zn3V4(PO4)6 can still be considered as ionic material.

3.2. Intrinsic Defect Properties

Here, we calculated the Frenkel, Schottky, and anti-site defect formation energy processes. These defects are crucial as they determine the properties of materials. As point defect (vacancy and interstitial) energies are necessary to calculate the defect process energies, vacancy and interstitial formation energies were first calculated. The defect processes are written using Kröger–Vink notations [58]. In a Schottky defect process, vacancies are introduced in the lattice. A Frenkel defect consists of a vacancy–interstitial pair. The concentration of these two defect processes is temperature-dependent. In general, the concentration of these defects will increase with increasing temperature. The Frenkel defect process controls the ionic conductivity of a material as explained in a previous experimental study on AgI [59]. An enhancement in the electronic conductivity was explained in the presence of Schottky defects in LaFeO3 [60]. Point defect energies were combined to calculate Frenkel, Schottky, and anti-site defect formation energies.
Zn   Frenkel :   Zn Zn X     V Zn +   Zn i
V   Frenkel :   V V X     V V +   V i
P   Frenkel :   P P X     V P +   P i
O   Frenkel :   O O X     V O +   O i
Schottky : 3   Zn Zn   X + 4   V V X   + 6   P P X + 24   O O X 3   V Zn + 4   V V + 6   V P + 12   V O +   Zn 3 V 4 ( PO 4 ) 6
ZnO   Schottky :   Zn Zn X +   O O X     V Zn + V O + ZnO
V 2 O 3   Schottky :   2   V V X + 3   O O X     2   V V + 3   V O + V 2 O 3
P 2 O 5   Schottky :   2   P P X + 5   O O X     2   V P + 5   V O + P 2 O 5
Zn / V   anti site   isolated :   Zn Zn X +   V V X   Zn V + V Zn
Zn / V   anti site   cluster :   Zn Zn X +   V V X   { Zn V + V Zn } X  
In Figure 2, calculated defect energies are reported. The Zn-V anti-site cluster exhibits the lowest defect formation energy. In this defect process, a small amount of Zn2+ ions will be on the V site and vice versa. The defect formation energy difference between the anti-site (isolated) and anti-site (cluster) is defined as the binding energy (−0.18 eV). Exoergic binding shows that isolated defects ( Zn V   and   V Zn ) will aggregate as soon as the isolated defects are present in this material. This defect has been determined in a variety of oxide materials both experimentally and theoretically [61,62,63,64,65]. The influence of the cation anti-site defect has been discussed in an experimental study on LiFePO4 [66]. The electrochemical performance of LiNiPO4 upon the Li-Ni anti-site defect has been discussed by Kempaiah et al. [67]. The cycled structure of Li2FeSiO4 exhibited Li-Fe ion mixing and the Li-ion diffusion pathways and activation energies were different to those found in its as-prepared structure [68]. In a recent study, defect and diffusion properties of spinel-ZnCo2O4 were examined using classical simulation [40]. It was found that the Zn-Co anti-site cluster is the lowest-energy defect with a binding energy of −0.14 eV.
Partial Schottky (ZnO and V2O3) and Zn Frenkel defect energies are closer to each other. The Zn Frenkel is an important defect energy process as it can govern the vacancy-assisted Zn-ion migration in this material. This defect energy is 2.82 eV, much lower than that calculated in ZnCo2O4. The O Frenkel has a defect energy of 3.00 eV. This is higher only by ~0.30 eV than that calculated in the partial Schottky or Zn Frenkel. Other Schottky or Frenkel defects are higher-defect-energy processes and do not form under normal temperatures. In particular, the P Frenkel has the highest defect energy of 10.60 eV.

3.3. Diffusion of Zn2+ Ions

The understanding of Zn2+ ion diffusion is important as it determines the overall performance of Zn3V4(PO4)6. Here, we used classical simulation techniques to calculate the energies of activation of local Zn hops and construct diffusion pathways with long range. In a previous classical simulation study [69], it was reported that Li+ ion migration in LiFePO4 is one-dimensional with a curve that was later confirmed in a neutron diffraction study [70].
Five local Zn hops were found. In the first hop (A), Zn diffuses in the bc plane having an activation energy of 1.05 eV (see Figure 3). A jump distance of 3.98 Å and its activation energy 1.88 eV were calculated in the second hop. Both activation energies are quite high and, therefore, the diffusion of Zn2+ would be slow. Table 4 lists the activation energies of each hoping distance. In particular, longer hop distances (>4.00 Å) have higher activation energies (>3 eV). This means that Zn2+ ions would diffuse very slowly via those hops. The slow diffusivity of Zn2+ can be due to several reasons including the high positive charge of 2+, long Zn hop distance, and the crystal structure. Morkhova et al. [71] recently employed DFT simulations to calculate activation energies of some Zn2+ ion conductors. The most promising structures were found to be spinel compounds with the chemical formula of ZnM2O4 (M = Fe, Co, Cr, and V) and their activation energies ranged between 0.54 eV and 0.68 eV. In the crystal structure of Zn3S2O9, the activation energy of Zn2+ is high (1.55 eV), although this compound consists of three Zn2+ ions per formula unit. In a previous classical simulation study of spinel-ZnCo2O4, it was reported that the Zn2+ ions migrate in a linear pathway having an activation energy of 0.71 eV [40].
An improvement of Zn2+ ion diffusion can be made by preparing materials at the nanoscale, lowering the Zn-Zn separation. The polymerization intercalation method has also been applied to enhance the rate of diffusion of Zn2+ ions [72]. In this method, the electrostatic interaction between Zn2+ and O2− ions is weakened to overcome the sluggish diffusion of Zn2+ ions. A combined experimental and DFT study of ZnCo2O4 showed that the oxygen vacancy formation can improve the Zn2+ ion diffusion by enlarging channels [73]. A facile hydrothermal method was applied to prepare a composite consisting of ZnxV2O5 and graphene oxide to form a stabilized structure and enhance the diffusion of Zn2+ ions [74].

3.4. Solution of Dopants

Substitutional doping of elements is an important process to modify the properties of a material [75]. The ionic conductivity of zirconia was enhanced by the doping of yttria [76]. Such an enhancement was shown to associate with the point defects controlling the oxygen diffusion at grain boundaries. Ru doping on the Fe site reduced the Li-Li hop distances and enhanced the Li-ion diffusion in LiFePO4 [77]. Here, we considered alkali earth (Mg, Ca, Sr, and Ba), divalent transition metal (Co, Mn, Fe, and Ni), and trivalent (Al, Ga, Gd, In, Sc, Y, and La) dopants to predict candidate dopants that can be tried experimentally. Appropriate lattice energies of dopant oxides and charge-compensating defects were introduced to construct defect reaction equations. Table 5 reports the potential parameters used for dopant oxides in this study.
First, alkali earth dopants (R = Mg, Ca, Sr, and Ba) were substitutionally doped on the Zn site. The following equation explains the doping process.
RO   +   Zn Zn X   R Zn X +   ZnO
Solution energies are reported in Figure 4a. The Ca2+ is found to be the most favorable dopant. A negative solution energy (−0.16 eV) is calculated for this dopant. The solution energy calculated for the Mg2+ is also negative but lower by 0.05 eV than that calculated for the Ca2+. These dopants are favored partly due to their ionic radii (Mg2+: 0.89 Å and Ca2+: 1.12 Å) closely matching with the ionic radius of Zn2+ (0.88 Å). The solution of Sr2+ is endoergic with the solution energy of 0.41 eV. The Ba2+ exhibits a high positive solution energy of 1.73 eV, meaning that this dopant can be doped only at high temperatures. The total DOS plot shows that the Ca-doped Zn3V4(PO4)6 is still metallic (see Figure 4b). The states of the Ca lie in the deeper level of the valence band (see Figure 4c).
Next, divalent transition metal ions (Ni2+, Co2+, Fe2+, and Mn2+) were considered on the Zn site. The most promising dopant is Fe2+ and its solution energy is –1.30 eV (see Figure 5a). The negative solution energy indicates that the Fe2+ on the Zn site is thermodynamically stable. The second most favorable dopant is Mn2+. The solution energy for this dopant is negative (−0.08 eV), meaning that this dopant is also promising. Both Ni2+ and Co2+ exhibit positive solution energies. The most unfavorable dopant is Ni2+. The metallic characteristic of Zn3V4(PO4)6 is retained upon Fe-doping (see Figure 5b). The states in the Fermi energy level are significantly affected by the d-states of Fe (see Figure 5c).
Next, trivalent dopants were substituted on the V site. This doping process, as other processes mentioned above, requires no charge-compensating defects as defined by the following equation.
R 2 O 3 + 2   V V X   2   R V X + V 2 O 3
The solution energies reported in Figure 6a indicate that most of the dopants except Al3+ exhibit an exoergic solution. The most promising dopant of In3+ has the solution energy of −1.10 eV. The least promising dopant of Al3+ exhibits an endothermic solution energy of 0.30 eV. There is an increase in solution energy with increasing ionic radius from Al3+ to In3+. Then, there is a gradual drop in the solution energy up to Ga3+. The solution energy of La3+ becomes more negative than that calculated for Gd3+.
The total DOS plot shows that In-doped Zn3V4(PO4)6 is metallic (see Figure 6b) and the states occupied at the Fermi energy level are a mixture of s, p, and d states of In (see Figure 6c).
A deformation charge density plot associated with the most favorable dopants interacting the lattice structure is provided in Figure 7.

3.5. Synthetic Routes for the Formation of Zn3V4(PO4)6

The synthesis of Zn3V4(PO4)6 was carried out using a high-temperature roasting method. In this complex method, CH3COO)2Zn·2H2O, NH4VO3, and H2C2O4·2H2O are used as starting materials [39]. Here, we consider some chemical reactions and calculate the formation energies by optimizing different oxides containing Zn, V, and P. In all cases, the formation of one mole of Zn3V4(PO4)6 is considered and the reaction energies are exothermic (see Table 6). This means all five routes are theoretically feasible. In the first reaction, the reaction liberates energy (−6.27 eV) in the form of heat. In this reaction route, binary oxides are taken as reactants. The second reaction has two binary oxides and one ternary oxide. The inclusion of a ternary oxide results in a lower reaction energy of −3.12 eV. In the third reaction, a ternary oxide consisting of Zn, P, and O (Zn2P2O7) is considered as a reactant. This reaction is still exothermic and its reaction energy is lower than those of reactions 2 and 3. In the least feasible reaction, a ternary oxide consisting of high Zn content (Zn3(PO4)2) is used. This reaction is exothermic with a reaction energy of −0.29 eV. Although all the reaction energies are thermodynamically exoergic, the availability, abundance, cost of materials, and ease of experiments should also be considered. In addition, kinetic feasibility is also important as the start of the reaction depends on the activation energy. The kinetic barrier can be dealt with by using appropriate catalysts.

4. Conclusions

In conclusion, we have studied the nature of defects, migration of Zn2+ ions, a solution of isovalent dopants, and the thermodynamic stability of different reaction routes for the synthesis of Zn3V4(PO4)6 using computational modeling techniques. The Zn-V anti-site is the most energetically favorable defect process. The diffusion of Zn2+ ions is low in this material. The Ca2+ and Fe2+ ions are the most promising dopant ions on the Zn site, while the V site prefers the In3+ dopant ion. The formation of Zn3V4(PO4)6 is favored by the reaction between binary oxides (ZnO, V2O3, and P2O5).

Funding

This research received no external funding.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The high-performance computing center at Imperial College London is acknowledged for providing computational facilities to run VASP code.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Goodenough, J.B.; Park, K.-S. The Li-Ion Rechargeable Battery: A Perspective. J. Am. Chem. Soc. 2013, 135, 1167–1176. [Google Scholar] [CrossRef] [PubMed]
  2. Wang, W.; Wang, J.; Tian, J.; Lu, J.; Xiong, R. Application of Digital Twin in Smart Battery Management Systems. Chin. J. Mech. Eng. 2021, 34, 57. [Google Scholar] [CrossRef]
  3. Sun, Y.-K. Promising All-Solid-State Batteries for Future Electric Vehicles. ACS Energy Lett. 2020, 5, 3221–3223. [Google Scholar] [CrossRef]
  4. Manthiram, A. An Outlook on Lithium Ion Battery Technology. ACS Central Sci. 2017, 3, 1063–1069. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Kim, T.; Song, W.; Son, D.-Y.; Ono, L.K.; Qi, Y. Lithium-ion batteries: Outlook on present, future, and hybridized technologies. J. Mater. Chem. A 2019, 7, 2942–2964. [Google Scholar] [CrossRef]
  6. Grey, C.P.; Hall, D.S. Prospects for lithium-ion batteries and beyond—A 2030 vision. Nat. Commun. 2020, 11, 6279. [Google Scholar] [CrossRef]
  7. Nitta, N.; Wu, F.; Lee, J.T.; Yushin, G. Li-ion battery materials: Present and future. Mater. Today 2014, 18, 252–264. [Google Scholar] [CrossRef]
  8. Blomgren, G.E. The Development and Future of Lithium Ion Batteries. J. Electrochem. Soc. 2016, 164, A5019–A5025. [Google Scholar] [CrossRef] [Green Version]
  9. Edge, J.S.; O’Kane, S.; Prosser, R.; Kirkaldy, N.D.; Patel, A.N.; Hales, A.; Ghosh, A.; Ai, W.; Chen, J.; Yang, J.; et al. Lithium ion battery degradation: What you need to know. Phys. Chem. Chem. Phys. 2021, 23, 8200–8221. [Google Scholar] [CrossRef]
  10. Palomares, V.; Serras, P.; Villaluenga, I.; Hueso, K.B.; Carretero-González, J.; Rojo, T. Na-ion batteries, recent advances and present challenges to become low cost energy storage systems. Energy Environ. Sci. 2012, 5, 5884–5901. [Google Scholar] [CrossRef]
  11. Abraham, K.M. How Comparable Are Sodium-Ion Batteries to Lithium-Ion Counterparts? ACS Energy Lett. 2020, 5, 3544–3547. [Google Scholar] [CrossRef]
  12. Anoopkumar, V.; John, B.; Mercy, T.D. Potassium-Ion Batteries: Key to Future Large-Scale Energy Storage? ACS Appl. Energy Mater. 2020, 3, 9478–9492. [Google Scholar]
  13. Min, X.; Xiao, J.; Fang, M.; Wang, W.; Zhao, Y.; Liu, Y.; Abdelkader, A.M.; Xi, K.; Kumar, R.V.; Huang, Z. Potassium-ion batteries: Outlook on present and future technologies. Energy Environ. Sci. 2021, 14, 2186–2243. [Google Scholar] [CrossRef]
  14. You, C.; Wu, X.; Yuan, X.; Chen, Y.; Liu, L.; Zhu, Y.; Fu, L.; Wu, Y.; Guo, Y.-G.; van Ree, T. Advances in rechargeable Mg batteries. J. Mater. Chem. A 2020, 8, 25601–25625. [Google Scholar] [CrossRef]
  15. Lu, D.; Liu, H.; Huang, T.; Xu, Z.; Ma, L.; Yang, P.; Qiang, P.; Zhang, F.; Wu, D. Magnesium ion based organic secondary batteries. J. Mater. Chem. A 2018, 6, 17297–17302. [Google Scholar] [CrossRef]
  16. Hosein, I.D. The Promise of Calcium Batteries: Open Perspectives and Fair Comparisons. ACS Energy Lett. 2021, 6, 1560–1565. [Google Scholar] [CrossRef]
  17. Gummow, R.J.; Vamvounis, G.; Kannan, M.B.; He, Y. Calcium-Ion Batteries: Current State-of-the-Art and Future Perspectives. Adv. Mater. 2018, 30, e1801702. [Google Scholar] [CrossRef]
  18. Li, G.; Huang, Z.; Chen, J.; Yao, F.; Liu, J.; Li, O.L.; Sun, S.; Shi, Z. Rechargeable Zn-ion batteries with high power and energy densities: A two-electron reaction pathway in birnessite MnO2 cathode materials. J. Mater. Chem. A 2019, 8, 1975–1985. [Google Scholar] [CrossRef]
  19. Fang, G.; Zhou, J.; Pan, A.; Liang, S. Recent Advances in Aqueous Zinc-Ion Batteries. ACS Energy Lett. 2018, 3, 2480–2501. [Google Scholar] [CrossRef]
  20. Zampardi, G.; La Mantia, F. Open challenges and good experimental practices in the research field of aqueous Zn-ion batteries. Nat. Commun. 2022, 13, 687. [Google Scholar] [CrossRef]
  21. Mallick, S.; Raj, C.R. Aqueous Rechargeable Zn-ion Batteries: Strategies for Improving the Energy Storage Performance. Chemsuschem 2021, 14, 1987–2022. [Google Scholar] [CrossRef]
  22. Blanc, L.E.; Kundu, D.; Nazar, L.F. Scientific Challenges for the Implementation of Zn-Ion Batteries. Joule 2020, 4, 771–799. [Google Scholar] [CrossRef]
  23. Lahiri, A.; Yang, L.; Li, G.; Endres, F. Mechanism of Zn-Ion Intercalation/Deintercalation in a Zn–Polypyrrole Secondary Battery in Aqueous and Bio-Ionic liquid Electrolytes. ACS Appl. Mater. Interfaces 2019, 11, 45098–45107. [Google Scholar] [CrossRef] [PubMed]
  24. Xu, S.; Sun, M.; Wang, Q.; Wang, C. Recent progress in organic electrodes for zinc-ion batteries. J. Semicond. 2020, 41, 091704. [Google Scholar] [CrossRef]
  25. Du, H.; Zheng, Y.; Dou, Z.; Zhan, H. Zn-Doped LiNi1/3Co1/3Mn1/3O2 Composite as Cathode Material for Lithium Ion Battery: Preparation, Characterization, and Electrochemical Properties. J. Nanomater. 2015, 2015, 867618. [Google Scholar] [CrossRef] [Green Version]
  26. Wang, L.; Huang, K.-W.; Chen, J.; Zheng, J. Ultralong cycle stability of aqueous zinc-ion batteries with zinc vanadium oxide cathodes. Sci. Adv. 2019, 5, eaax4279. [Google Scholar] [CrossRef]
  27. Shen, H.; Liu, B.; Nie, Z.; Li, Z.; Jin, S.; Huang, Y.; Zhou, H. A comparison study of MnO2 and Mn2O3 as zinc-ion battery cathodes: An experimental and computational investigation. RSC Adv. 2021, 11, 14408–14414. [Google Scholar] [CrossRef]
  28. Zhang, Y.; Liu, Y.; Liu, Z.; Wu, X.; Wen, Y.; Chen, H.; Ni, X.; Liu, G.; Huang, J.; Peng, S. MnO2 cathode materials with the improved stability via nitrogen doping for aqueous zinc-ion batteries. J. Energy Chem. 2021, 64, 23–32. [Google Scholar] [CrossRef]
  29. Venkatesan, R.; Bauri, R.; Mayuranathan, K.K. Zinc Vanadium Oxide Nanobelts as High-Performance Cathodes for Rechargeable Zinc-Ion Batteries. Energy Fuels 2022, 36, 7854–7864. [Google Scholar] [CrossRef]
  30. Li, Y.; Zhao, J.; Hu, Q.; Hao, T.; Cao, H.; Huang, X.; Liu, Y.; Zhang, Y.; Lin, D.; Tang, Y.; et al. Prussian blue analogs cathodes for aqueous zinc ion batteries. Mater. Today Energy 2022, 29, 101095. [Google Scholar] [CrossRef]
  31. Cai, K.; Luo, S.; Feng, J.; Wang, J.; Zhan, Y.; Wang, Q.; Zhang, Y.; Liu, X. Recent Advances on Spinel Zinc Manganate Cathode Materials for Zinc-Ion Batteries. Chem. Rec. 2021, 22, e202100169. [Google Scholar] [CrossRef] [PubMed]
  32. Kundu, D.; Oberholzer, P.; Glaros, C.; Bouzid, A.; Tervoort, E.; Pasquarello, A.; Niederberger, M. Organic Cathode for Aqueous Zn-Ion Batteries: Taming a Unique Phase Evolution toward Stable Electrochemical Cycling. Chem. Mater. 2018, 30, 3874–3881. [Google Scholar] [CrossRef]
  33. Li, X.; Chen, Z.; Yang, Y.; Liang, S.; Lu, B.; Zhou, J. The phosphate cathodes for aqueous zinc-ion batteries. Inorg. Chem. Front. 2022, 9, 3986–3998. [Google Scholar] [CrossRef]
  34. Li, G.; Yang, Z.; Jiang, Y.; Jin, C.; Huang, W.; Ding, X.; Huang, Y. Towards polyvalent ion batteries: A zinc-ion battery based on NASICON structured Na3V2(PO4)3. Nano Energy 2016, 25, 211–217. [Google Scholar] [CrossRef]
  35. Zhao, H.B.; Hu, C.J.; Cheng, H.; Fang, J.H.; Xie, Y.P.; Fang, W.Y.; Doan, T.N.L.; Hoang, T.K.A.; Xu, J.Q.; Chen, P. Novel Rechargeable M3V2(PO4)3//Zinc (M = Li, Na) Hybrid Aqueous Batteries with Excellent Cycling Performance. Sci. Rep. 2016, 6, 25809. [Google Scholar] [CrossRef] [Green Version]
  36. Li, W.; Wang, K.; Cheng, S.; Jiang, K. A long-life aqueous Zn-ion battery based on Na3V2(PO4)2F3 cathode. Energy Storage Mater. 2018, 15, 14–21. [Google Scholar] [CrossRef]
  37. Yesibolati, N.; Umirov, N.; Koishybay, A.; Omarova, M.; Kurmanbayeva, I.; Zhang, Y.; Zhao, Y.; Bakenov, Z. High Performance Zn/LiFePO4 Aqueous Rechargeable Battery for Large Scale Applications. Electrochim. Acta 2015, 152, 505–511. [Google Scholar] [CrossRef] [Green Version]
  38. Verma, V.; Kumar, S.; Manalastas, W.J.; Zhao, J.; Chua, R.; Meng, S.; Kidkhunthod, P.; Srinivasan, M. Layered VOPO4 as a Cathode Material for Rechargeable Zinc-Ion Battery: Effect of Polypyrrole Intercalation in the Host and Water Concentration in the Electrolyte. ACS Appl. Energy Mater. 2019, 2, 8667–8674. [Google Scholar] [CrossRef]
  39. Zhao, Q.; Zhu, Y.; Liu, S.; Liu, Y.; He, T.; Jiang, X.; Yang, X.; Feng, K.; Hu, J. Zn3V4(PO4)6: A New Rocking-Chair-Type Cathode Material with High Specific Capacity Derived from Zn2+/H+ Cointercalation for Aqueous Zn-Ion Batteries. ACS Appl. Mater. Interfaces 2022, 14, 32066–32074. [Google Scholar] [CrossRef]
  40. Morando, C.; Cofrancesco, P.; Tealdi, C. Zn ion diffusion in spinel-type cathode materials for rechargeable batteries: The role of point defects. Mater. Today Commun. 2020, 25, 101478. [Google Scholar] [CrossRef]
  41. Fisher, C.A.J.; Prieto, V.M.H.; Islam, M.S. Lithium Battery Materials LiMPO4 (M = Mn, Fe, Co, and Ni): Insights into Defect Association, Transport Mechanisms, and Doping Behavior. Chem. Mater. 2008, 20, 5907–5915. [Google Scholar] [CrossRef]
  42. Kuganathan, N.; Chroneos, A. Na3V(PO4)2 cathode material for Na ion batteries: Defects, dopants and Na diffusion. Solid State Ion. 2019, 336, 75–79. [Google Scholar] [CrossRef] [Green Version]
  43. Kuganathan, N.; Chroneos, A. Defect Chemistry and Na-Ion Diffusion in Na3Fe2(PO4)3 Cathode Material. Materials 2019, 12, 1348. [Google Scholar] [CrossRef] [Green Version]
  44. Kuganathan, N.; Rushton, M.J.D.; Grimes, R.W.; Kilner, J.A.; Gkanas, E.I.; Chroneos, A. Self-diffusion in garnet-type Li7La3Zr2O12 solid electrolytes. Sci. Rep. 2021, 11, 451. [Google Scholar] [CrossRef]
  45. Reischl, B.R.; Raiteri, P.; Gale, J.D.; Rohl, A.L. Atomistic Simulation of Atomic Force Microscopy Imaging of Hydration Layers on Calcite, Dolomite, and Magnesite Surfaces. J. Phys. Chem. C 2019, 123, 14985–14992. [Google Scholar] [CrossRef] [Green Version]
  46. Gale, J.D.; Rohl, A.L. The General Utility Lattice Program (GULP). Mol. Simul. 2003, 29, 291–341. [Google Scholar] [CrossRef]
  47. Lewis, G.V.; Catlow, C.R.A. Potential models for ionic oxides. J. Phys. C Solid State Phys. 1985, 18, 1149–1161. [Google Scholar] [CrossRef]
  48. Kuganathan, N.; Ganeshalingam, S.; Chroneos, A. Defects, Dopants and Lithium Mobility in Li9V3(P2O7)3(PO4)2. Sci. Rep. 2018, 8, 8140. [Google Scholar] [CrossRef] [PubMed]
  49. Gale, J.D. GULP: A computer program for the symmetry-adapted simulation of solids. J. Chem. Soc. Faraday Trans. 1997, 93, 629–637. [Google Scholar] [CrossRef]
  50. Mott, N.F.; Littleton, M.J. Conduction in polar crystals. I. Electrolytic conduction in solid salts. Trans. Faraday Soc. 1938, 34, 485–499. [Google Scholar] [CrossRef]
  51. 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]
  52. Blöchl, P.E. Projector augmented-wave method. Phys. Rev. B Condens. Matter Mater. Phys. 1994, 50, 17953–17979. [Google Scholar] [CrossRef] [Green Version]
  53. Monkhorst, H.J.; Pack, J.D. Special points for Brillouin-zone integrations. Phys. Rev. B 1976, 13, 5188. [Google Scholar] [CrossRef]
  54. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Press, W.H.; Teukolsky, S.A.; Vetterling, W.T.; Flannery, B.P. Numerical Recipes in C: The Art of Scientific Computing, 2nd ed.; Cambridge University Press: Cambridge, UK, 1992. [Google Scholar]
  56. Henkelman, G.; Arnaldsson, A.; Jónsson, H. A fast and robust algorithm for Bader decomposition of charge density. Comput. Mater. Sci. 2005, 36, 354–360. [Google Scholar] [CrossRef]
  57. Boudin, S.; Grandin, A.; LeClaire, A.; Borel, M.; Raveau, B. The Original Structure of Zn3V4(PO4)6 Involving Bioctahedral V2O10 Units and ZnO5 Trigonal Bipyramids. J. Solid State Chem. 1995, 115, 140–145. [Google Scholar] [CrossRef]
  58. Kröger, F.A.; Vink, H.J. Relations between the Concentrations of Imperfections in Crystalline Solids. In Solid State Physics; Seitz, F., Turnbull, D., Eds.; Academic Press: Cambridge, MA, USA, 1956; Volume 3, pp. 307–435. [Google Scholar]
  59. Lee, J.S.; Adams, S.; Maier, J. Defect chemistry and transport characteristics of β-AgI. J. Phys. Chem. Solids 2000, 61, 1607–1622. [Google Scholar] [CrossRef]
  60. Wærnhus, I.; Grande, T.; Wiik, K. Electronic properties of polycrystalline LaFeO3. Part II: Defect modelling including Schottky defects. Solid State Ion. 2005, 176, 2609–2616. [Google Scholar] [CrossRef]
  61. Chen, H.; Millis, A. Antisite defects at oxide interfaces. Phys. Rev. B 2016, 93, 104111. [Google Scholar] [CrossRef] [Green Version]
  62. Saloaro, M.; Liedke, M.O.; Angervo, I.; Butterling, M.; Hirschmann, E.; Wagner, A.; Huhtinen, H.; Paturi, P. Exploring the anti-site disorder and oxygen vacancies in Sr2FeMoO6 thin films. J. Magn. Magn. Mater. 2021, 540, 168454. [Google Scholar] [CrossRef]
  63. Dey, U.; Chatterjee, S.; Taraphder, A. Antisite-disorder engineering in La-based oxide heterostructures via oxygen vacancy control. Phys. Chem. Chem. Phys. 2018, 20, 17871–17880. [Google Scholar] [CrossRef] [PubMed]
  64. Lee, Y.-L.; Holber, J.; Paudel, H.P.; Sorescu, D.C.; Senor, D.J.; Duan, Y. Density functional theory study of the point defect energetics in γ-LiAlO2, Li2ZrO3 and Li2TiO3 materials. J. Nucl. Mater. 2018, 511, 375–389. [Google Scholar] [CrossRef]
  65. Oba, F.; Choi, M.; Togo, A.; Tanaka, I. Point defects in ZnO: An approach from first principles. Sci. Technol. Adv. Mater. 2011, 12, 034302. [Google Scholar] [CrossRef] [PubMed]
  66. Liu, H.; Choe, M.-J.; Enrique, R.A.; Orvañanos, B.; Zhou, L.; Liu, T.; Thornton, K.; Grey, C.P. Effects of Antisite Defects on Li Diffusion in LiFePO4 Revealed by Li Isotope Exchange. J. Phys. Chem. C 2017, 121, 12025–12036. [Google Scholar] [CrossRef]
  67. Devaraju, M.K.; Truong, Q.D.; Hyodo, H.; Sasaki, Y.; Honma, I. Synthesis, characterization and observation of antisite defects in LiNiPO4 nanomaterials. Sci. Rep. 2015, 5, 11041. [Google Scholar] [CrossRef] [Green Version]
  68. Armstrong, A.R.; Kuganathan, N.; Islam, M.S.; Bruce, P.G. Structure and Lithium Transport Pathways in Li2FeSiO4 Cathodes for Lithium Batteries. J. Am. Chem. Soc. 2011, 133, 13031–13035. [Google Scholar] [CrossRef] [Green Version]
  69. Islam, M.S.; Driscoll, D.J.; Fisher, C.A.J.; Slater, P.R. Atomic-Scale Investigation of Defects, Dopants, and Lithium Transport in the LiFePO4 Olivine-Type Battery Material. Chem. Mater. 2005, 17, 5085–5092. [Google Scholar] [CrossRef]
  70. Nishimura, S.-I.; Kobayashi, G.; Ohoyama, K.; Kanno, R.; Yashima, M.; Yamada, A. Experimental visualization of lithium diffusion in LixFePO4. Nat. Mater. 2008, 7, 707–711. [Google Scholar] [CrossRef]
  71. Morkhova, Y.A.; Rothenberger, M.; Leisegang, T.; Adams, S.; Blatov, V.A.; Kabanov, A.A. Computational Search for Novel Zn-Ion Conductors—A Crystallochemical, Bond Valence, and Density Functional Study. J. Phys. Chem. C 2021, 125, 17590–17599. [Google Scholar] [CrossRef]
  72. Bin, D.; Huo, W.; Yuan, Y.; Huang, J.; Liu, Y.; Zhang, Y.; Dong, F.; Wang, Y.; Xia, Y. Organic-Inorganic-Induced Polymer Intercalation into Layered Composites for Aqueous Zinc-Ion Battery. Chem 2020, 6, 968–984. [Google Scholar] [CrossRef]
  73. Huang, J.; Li, Y.; Xie, R.; Li, J.; Tian, Z.; Chai, G.; Zhang, Y.; Lai, F.; He, G.; Liu, C.; et al. Structural engineering of cathodes for improved Zn-ion batteries. J. Energy Chem. 2021, 58, 147–155. [Google Scholar] [CrossRef]
  74. Fan, Y.; Yu, X.; Feng, Z.; Hu, M.; Zhang, Y. Synthesis of Zn2+-Pre-Intercalated V2O5·nH2O/rGO Composite with Boosted Electrochemical Properties for Aqueous Zn-Ion Batteries. Molecules 2022, 27, 5387. [Google Scholar] [CrossRef] [PubMed]
  75. Zhang, K.; Sheng, H.; Wu, X.-W.; Fu, L.; Liu, Z.; Zhou, C.; Holze, R.; Wu, Y. Improving electrochemical properties by sodium doping for lithium-rich layered oxides. ACS Appl. Energy Mater. 2020, 3, 8953–8959. [Google Scholar] [CrossRef]
  76. Park, J.S.; Kim, Y.-B.; An, J.; Prinz, F.B. Oxygen diffusion across the grain boundary in bicrystal yttria stabilized zirconia. Solid State Commun. 2012, 152, 2169–2171. [Google Scholar] [CrossRef]
  77. Gao, Y.; Xiong, K.; Zhang, H.; Zhu, B. Effect of Ru Doping on the Properties of LiFePO4/C Cathode Materials for Lithium-Ion Batteries. ACS Omega 2021, 6, 14122–14129. [Google Scholar] [CrossRef]
Figure 1. (a) Crystal structure of Zn3V4(PO4)6, (b) total DOS plot, and (c) charge density plot showing the bonding nature in Zn3V4(PO4)6. The vertical blue dotted lines correspond to the Fermi energy level.
Figure 1. (a) Crystal structure of Zn3V4(PO4)6, (b) total DOS plot, and (c) charge density plot showing the bonding nature in Zn3V4(PO4)6. The vertical blue dotted lines correspond to the Fermi energy level.
Batteries 09 00005 g001
Figure 2. Energies of defects for different defect processes in Zn3V4(PO4)6.
Figure 2. Energies of defects for different defect processes in Zn3V4(PO4)6.
Batteries 09 00005 g002
Figure 3. (a) Local Zn hops considered, (b) the most possible long-range diffusion pathway consisting of local hops A and B, (c) energy profile diagram showing the activation energy of the local hop A, and (d) similar diagram plotted for the local hop B.
Figure 3. (a) Local Zn hops considered, (b) the most possible long-range diffusion pathway consisting of local hops A and B, (c) energy profile diagram showing the activation energy of the local hop A, and (d) similar diagram plotted for the local hop B.
Batteries 09 00005 g003
Figure 4. (a) Solution energies of alkali earth dopants on the Zn site, (b) total DOS plot of Ca-substituted Zn3V4(PO4)6, and (c) DOS plot associated with the Ca.
Figure 4. (a) Solution energies of alkali earth dopants on the Zn site, (b) total DOS plot of Ca-substituted Zn3V4(PO4)6, and (c) DOS plot associated with the Ca.
Batteries 09 00005 g004
Figure 5. (a) Energies of solution calculated for transition metal dopants on the Zn site, (b) total DOS plot of Fe-substituted Zn3V4(PO4)6, and (c) DOS plot associated with the Fe.
Figure 5. (a) Energies of solution calculated for transition metal dopants on the Zn site, (b) total DOS plot of Fe-substituted Zn3V4(PO4)6, and (c) DOS plot associated with the Fe.
Batteries 09 00005 g005
Figure 6. (a) Solution energies of trivalent dopants on the V site, (b) total DOS plot of In-substituted Zn3V4(PO4)6, and (c) DOS plot associated with the In.
Figure 6. (a) Solution energies of trivalent dopants on the V site, (b) total DOS plot of In-substituted Zn3V4(PO4)6, and (c) DOS plot associated with the In.
Batteries 09 00005 g006
Figure 7. Charge density plots associated with the most favorable dopants: (a) Ca2+, and (b) Fe3+ and (c) In3+.
Figure 7. Charge density plots associated with the most favorable dopants: (a) Ca2+, and (b) Fe3+ and (c) In3+.
Batteries 09 00005 g007
Table 1. Buckingham potential parameters used in the classical simulations of Zn3V4(PO4)6 [47,48]. Two-body (Φij (rij) = Aij exp (− rij/ρij) − Cij/rij6), where A, ρ, and C are parameters reproducing the experimental data. The values of Y and K are shell charges and spring constants, respectively.
Table 1. Buckingham potential parameters used in the classical simulations of Zn3V4(PO4)6 [47,48]. Two-body (Φij (rij) = Aij exp (− rij/ρij) − Cij/rij6), where A, ρ, and C are parameters reproducing the experimental data. The values of Y and K are shell charges and spring constants, respectively.
InteractionA/eVρ/ÅC/eV·Å6Y/eK/eV·Å–2
Zn2+ − O2−499.60.35950.002.0099,999
V3+ − O2−1410.820.311716.002.04196.3
P5+ − O2−897.26480.35770.005.00099,999
O2 − O222,764.300.149027.89−2.8074.92
Table 2. Calculated and experimental parameters of Zn3V4(PO4)6.
Table 2. Calculated and experimental parameters of Zn3V4(PO4)6.
ParameterCalculatedExperiment [57]|∆| (%)
ClassicalDFTClassicalDFT
a (Å)6.1726.3936.3492.790.69
b (Å)7.9607.9757.8691.161.35
c (Å)9.2089.3919.3241.240.78
α (°)105.43105.45105.320.100.12
β (°)108.31107.20108.660.331.34
γ (°)102.35101.97101.231.110.73
Table 3. Calculated Bader charges on the atoms in bulk Zn3V4(PO4)6.
Table 3. Calculated Bader charges on the atoms in bulk Zn3V4(PO4)6.
AtomBader Charge (e)
Zn+1.43
V+2.10
P+5.00
O−1.78
Table 4. Calculated activation energies of different Zn hops.
Table 4. Calculated activation energies of different Zn hops.
HopDistance (Å)Activation Energy (eV)
A3.551.05
B3.981.88
C5.873.48
D5.943.62
E6.243.80
Table 5. Interatomic potential parameters used for dopants in the atomistic simulations of Zn3V4(PO4)6. Two-body (Φij (rij) = Aij exp (−rij/ρij) − Cij/rij6).
Table 5. Interatomic potential parameters used for dopants in the atomistic simulations of Zn3V4(PO4)6. Two-body (Φij (rij) = Aij exp (−rij/ρij) − Cij/rij6).
InteractionA (eV)ρ (Å)C (eV·Å6)Y (e)K (eV·Å−2)
Mg+ − O2−946.6270.318130.0002.00099,999
Ca2+ − O2−1090.400.33720.0001.26034.00
Sr2+ − O2−776.840.358670.0001.52611.406
Ba2+ − O2−931.790.39490.0001.46014.78
Ni2+ − O2−683.50.33320.0002.0008.77
Co2+ − O2−696.30.33620.0002.00010.74
Fe2+ − O2−1207.60.30840.0002.00099,999
Mn2+ − O2−715.800.34640.0003.00081.20
Al3+ − O2−1725.200.289710.0003.00099,999
Ga3+ − O2−2901.120.27420.0003.00099,999
Sc3+ − O2−1575.850.32110.0003.00099,999
In3+ − O2−1495.650.33274.333.00099,999
Y3+ − O2−1766.400.3384919.433.00099,999
Gd3+ − O2−1885.750.339920.343.00099,999
La3+ − O2−2088.790.346023.253.00099,999
Table 6. Different reaction routes for the synthesis of Zn3V4(PO4)6.
Table 6. Different reaction routes for the synthesis of Zn3V4(PO4)6.
Reaction NumberReactionReaction Energy (eV)
13 ZnO + 2 V2O3 + 3P2O5 → Zn3V4(PO4)6−6.27
23 ZnO + 4 VPO4 + P2O5 → Zn3V4(PO4)6−3.12
31.5 Zn2(P2O7) + 2 V2O3 + 1.5 P2O5 → Zn3V4(PO4)6−2.30
43 ZnO + V4(P2O7)3 → Zn3V4(PO4)6−1.72
53 Zn(PO3)2 + 2 V2O3 → Zn3V4(PO4)6−1.10
6Zn3(PO4)2 + 4 VPO4 → Zn3V4(PO4)6−0.29
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

Kuganathan, N. Defect Chemistry in Zn3V4(PO4)6. Batteries 2023, 9, 5. https://doi.org/10.3390/batteries9010005

AMA Style

Kuganathan N. Defect Chemistry in Zn3V4(PO4)6. Batteries. 2023; 9(1):5. https://doi.org/10.3390/batteries9010005

Chicago/Turabian Style

Kuganathan, Navaratnarajah. 2023. "Defect Chemistry in Zn3V4(PO4)6" Batteries 9, no. 1: 5. https://doi.org/10.3390/batteries9010005

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