Next Article in Journal
Early Stages of Crack Nucleation Mechanism in Fe39Mn20Co20Cr15Si5Al1 High-Entropy Alloy during Stress Corrosion Cracking Phenomenon: Pit Initiation and Growth
Previous Article in Journal
Shape-Memory Effect of 4D-Printed Gamma-Irradiated Low-Density Polyethylene
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Atomistic Simulation Studies of Na4SiO4

by
Mallikage Shalani Shanika
1,
Poobalasingam Abiman
1,
Poobalasuntharam Iyngaran
1 and
Navaratnarajah Kuganathan
2,*
1
Department of Chemistry, University of Jaffna, Thirunelvely, Jaffna 40000, Sri Lanka
2
Department of Materials, Faculty of Engineering, Imperial College London, London SW7 2AZ, UK
*
Author to whom correspondence should be addressed.
Crystals 2024, 14(8), 718; https://doi.org/10.3390/cryst14080718 (registering DOI)
Submission received: 24 July 2024 / Revised: 5 August 2024 / Accepted: 7 August 2024 / Published: 10 August 2024
(This article belongs to the Section Materials for Energy Applications)

Abstract

:
Tetrasodium silicate (Na4SiO4) has emerged as a promising candidate for battery applications due to its favorable ionic transport properties. Atomic-scale simulations employing classical pair potentials have elucidated the defect mechanisms and ion migration dynamics in Na4SiO4. The Na Frenkel defect, characterized by the creation of a Na vacancy and an interstitial Na⁺ ion, is identified as the most energetically favorable defect process, facilitating efficient vacancy-assisted Na⁺ ion migration. This process results in three-dimensional ion diffusion with a low activation energy of 0.55 eV, indicating rapid ion movement within the material. Among monovalent dopants (Li⁺, K⁺, and Rb⁺), K⁺ was found to be the most advantageous for substitution on the Na site. For trivalent doping, Al is the most favorable on the Si site, generating additional Na⁺ ions and potentially enhancing ionic conductivity. Ge was identified as a promising isovalent dopant for the Si site. These theoretical findings suggest that Na4SiO4 could offer high ionic conductivity and stability when optimized through appropriate doping. Experimental validation of these predictions could lead to the development of advanced battery materials with improved performance and durability.

1. Introduction

Creating clean, renewable energy is one of the major challenges of the 21st century, and overcoming this challenge necessitates the advancement of innovative material systems. Advances in materials science are crucial for enhancing the efficiency, cost-effectiveness, and sustainability of various energy technologies.
Batteries have transformed into a fundamental aspect of modern technology, seamlessly integrating into nearly every electronic device [1,2,3,4]. Acting as vital energy reservoirs, they have diverse applications across various industries [5,6,7]. Presently, lithium-ion batteries dominate the market, especially in electric vehicles and electronic gadgets [8,9,10]. However, the diminishing availability of lithium minerals poses a challenge to meeting demand. The world is currently grappling with a shortfall in lithium supply. In response, sodium-ion batteries have emerged as a focal point of extensive research, offering promise as a sustainable alternative to lithium-ion batteries [11,12,13,14,15]. Multiple molecules are currently under consideration as potential materials for anodes and cathodes in sodium-ion batteries [16,17,18].
The selection of electrode or electrolyte materials for Na-ion batteries is critical to achieving high performance, safety, and cost-effectiveness. Each material class has its advantages and challenges, and ongoing research focuses on optimizing these materials through structural engineering, composite formation, and surface modifications. Polyanionic-type materials are gaining significant attention in the development of Na-ion batteries due to their high working potential, excellent structural stability, and safety. These materials include compounds where the active cathode species are coordinated by anionic groups such as phosphates (e.g., NaMPO4 (M = Fe, Mn, Ni)), sulfates (e.g., Na2MSiO4 (M = Fe, Mn, Co, Ni)) and borates (e.g., Na3MB5O10 (M = V, Fe, Mn, Co, etc.)) [19,20,21,22].
Na4SiO4 holds particular significance in this pursuit [23]. As a mineral, it presents potential as a key component in the development of sodium-ion batteries. Research is actively investigating its potential and effectiveness in meeting the increasing demand for energy storage solutions [24,25,26]. Na4SiO4 can be produced through a wet mixing method using NaOH and Na2SiO3•9H2O as raw materials [27]. A key benefit of Na4SiO4 is its stability and eco-friendly characteristics. It can be easily prepared using simple, abundant raw materials at low cost. Furthermore, like orthophosphates, the orthosilicate (SiO44−) matrix benefits from strong Si–O bonds. These bonds contribute to the structural stability of the material, which is crucial for maintaining integrity during lithium insertion and extraction cycles. The study of CO2 chemisorption on Na4SiO4 doped with oxysalt surfaces using density functional theory (DFT) calculations highlights the potential of doping Na4SiO4 with oxysalts to create more effective materials for CO2 capture, contributing to efforts in mitigating climate change by reducing atmospheric CO2 levels [24]. The addition of alkali carbonates to Na4SiO4 enhances CO2 capture performance, especially at low temperatures, through the formation of C2O52− species. This has been confirmed by in situ Raman spectroscopy and supported by DFT calculations, providing a strong basis for developing improved CO2 capture technologies [28].
The study of defect processes in materials is a crucial aspect of materials science and engineering, as defects can significantly impact the physical, electrical, thermal, and mechanical properties of materials. Doping is a versatile and essential process in materials science, enabling the fine-tuning of material properties for a wide range of applications. Advances in doping techniques continue to drive innovation in electronics, optoelectronics, photovoltaics, and many other fields. Understanding and controlling doping processes is key to developing next-generation materials and devices. Na-ion diffusion is a complex but critical process in the development of Na-ion batteries and other sodium-based technologies. A deep understanding of the mechanisms, pathways, and influencing factors is essential for designing materials with optimal performance. Continued research, combining experimental studies with computational modeling, will be key to overcoming current challenges and advancing the field of Na-ion diffusion in materials.
To delve deeper into assessing the viability of Na4SiO4 as a battery material, it is crucial to thoroughly investigate additional properties such as defect energy, dopant energy, and ion migration. Conducting theoretical calculations can be challenging due to several factors including the complexity of the system and experimental validation. To overcome this obstacle, atomistic simulation studies have been previously utilized on {Kumar Prajapati, 2023 #6}to comprehensively ascertain and comprehend these properties [29,30,31,32,33,34].
In this study, we use atomistic simulation studies based on the classical pairwise potential to elucidate the defect, diffusion, and dopant properties of Na4SiO4.

2. Computational Methods

All calculations were performed using the classical pairwise potential simulation code GULP (General Utility Lattice Program) [35]. GULP is a computational tool used for modeling and simulating the properties of materials, particularly those with crystalline structures. It enables the calculation of various properties of materials, such as lattice dynamics, defect properties, and thermodynamic properties, using different interaction models and optimization algorithms. Interactions between ions in the crystal structure are modeled using long-range (Coulombic) and short-range (Pauli repulsion and van der Waals attraction) forces. Buckingham potentials (see Table 1) describe short-range repulsive forces. The BFGS algorithm (Broyden–Fletcher–Goldfarb–Shanno) is an iterative method for solving unconstrained nonlinear optimization problems. It is particularly well suited for problems where the function to be minimized (or maximized) is smooth. In the context of GULP, the BFGS algorithm is used to optimize the structure of a material by minimizing its potential energy [36]. In all relaxed configurations, the forces on all atoms are smaller than 0.001 eV/Å. The Mott–Littleton method [37] is used to model point defects and migrating ions. The Mott–Littleton method is a well-established technique used in the modeling of point defects and ion migration in crystalline materials. The Mott–Littleton method divides the crystal into two regions to model defects effectively. Region I (Inner Region) is the immediate vicinity around the defect where atomic positions are significantly displaced due to the presence of the defect. In this region, the interactions between atoms and the defect are calculated explicitly. Typically, this region is spherical, and the defect is located at its center. Region II (Outer Region) extends beyond Region I and is where the defect’s influence diminishes. The atoms in Region II are treated using continuum or semi-continuum approaches, where displacements are assumed to be small and can be described using linear elasticity theory or other approximations. Sodium-ion migration is calculated by considering seven interstitial points with equal intervals between neighboring Sodium sites. Defect energies of migrating ions at seven points along the diffusion path will be determined. The midpoint between two adjacent Na vacancy sites is the defect calculation center to reduce systematic errors. The energy difference between the maximum local energy associated with the saddle point along this diffusion path and the lowest Na vacancy formation energy will be calculated and reported as activation energy. Ions are modeled using this method as spherical objects with a full charge at the diluted limit. Defective energies are therefore likely to be overstated. However, relative energies will continue to trend in the same direction. The core-shell technique was used to model the polarization of ions. In earlier research [38,39], we described the formula for determining migratory paths and detailed activation energies of migration.

3. Results and Discussion

3.1. Modelling of Na4SiO4 Crystal Structure

Na4SiO4 shows a triclinic structure with the P-1 space group according to the X-ray diffraction pattern derived by Baur et al. [23] (CIF file name: ICSD_CollCode62594). This study further explains that it consists of 24 atoms (8 Na, 2 Si, and 14 O) in a unit cell (2 formula units) with the lattice parameter found in the x, y, and z directions as a = 5.58 Å, b = 5.58 Å, and c = 8.39 Å and α = 80.92°, β = 71.84°, and γ = 67.44°. Si atoms are coordinated with four oxygen atoms, forming a tetrahedral structure (see Figure 1). This tetrahedral coordination is a fundamental building block of many silicate structures, including quartz, feldspar, and various other minerals. To ensure the reliability of the classical potentials used in the study, a critical step is to validate these potentials against experimental data. This involves performing full geometry optimization of the crystal structure and comparing the calculated equilibrium lattice constants with experimentally determined values. The calculated lattice constants closely match the experimental values reported (see Table 2).

3.2. Defect Energetics

In this section, we examine the energetics of important defect processes in Na4SiO4. To investigate the defect characteristics in Na4SiO4, we initially analyzed point defects, including vacancies and interstitials, separately. These calculations form the basis for understanding more complex defect structures such as Schottky and Frenkel defects, as well as anti-site defects where Na and Si swap their atomic positions. The Schottky defect energy in Na4SiO4 was calculated by determining the individual vacancy formation energies for four Na⁺, one Si4+, and four O2− ions and summing them. This approach ensures that the defect maintains the stoichiometry and charge neutrality of the crystal. A Frenkel defect involves the displacement of an ion from its lattice site to an interstitial site, creating a vacancy–interstitial pair. The energy required to form a Na Frenkel defect is the sum of the vacancy formation energy and the interstitial formation energy for Na⁺. In Na4SiO4, anti-site defects involve Na and Si atoms swapping positions. For isolated anti-site defects, a single Na atom occupies a Si site, and a single Si atom occupies a Na site. For clustered anti-site defects, multiple pairs of Na and Si atoms swap positions. The following equations in Kröger–Vink notation [43] describe the Schottky, Frenkel, and anti-site defect processes.
N a   F r e n k e l : N a N a X   V N a + N a i
S i   F r e n k e l : S i S i X V S i + S i i
O   F r e n k e l : O O X V O + O i
S c h o t t k y : 4   N a N a X + S i S i X + 4   O O X 4   V N a + V S i + 4   V O + N a 4 S i O 4
N a 2 O   S c h o t t k y : 2   N a N a X + O O X 2   V N a + V O + N a 2 O
S i O 2   S c h o t t k y : S i S i X + 2   O O X V S i + 2   V O + S i O 2
N a / S i   a n t i s i t e ( i s o l a t e d ) : N a N a X + S i S i X N a S i + S i N a
N a / S i   a n t i s i t e ( c l u s t e r ) : N a N a X + S i S i X { N a S i + S i N a } X
Figure 2 demonstrates the individual defect reaction energies. The Na Frenkel defect has the lowest energy among the considered defect processes, with an energy of 1.34 eV per defect. This indicates that the Na ions in Na4SiO4 have a relatively low barrier for forming Frenkel defects, facilitating Na-ion migration via vacancy-assisted mechanisms. The next most energetically favorable defect process after the Na Frenkel defect is the Na2O Schottky defect (2.79 eV/defect). This defect involves the simultaneous creation of vacancies for two Na atoms and one O atom, effectively removing a Na2O unit from the crystal lattice. This higher energy implies that the formation of such defects, and consequently the loss of Na2O, is only feasible at high temperatures. Isolated anti-site defects have a high energy cost, making them unstable. Due to the exoergic nature of the clustering process (clustering energy: 4.91 − 8.77 = −3.86 eV), isolated anti-site defects tend to aggregate into clusters to achieve a lower energy state. This aggregation into clusters reflects a significant tendency for the system to favor clustered anti-site defects, which are more stable and energetically favorable compared to their isolated counterparts. Other Frenkel and Schottky defects have even higher formation energies, making them practically non-existent under typical conditions. The results indicate that Na Frenkel defects are the most energetically favorable and significant in Na4SiO4. Due to their lower formation energy, they dominate the defect landscape and influence the material’s behavior, particularly its ionic conductivity.

3.3. Na-Ion Migration

Examining sodium-ion diffusion pathways and activation energies is crucial for understanding the ionic conductivity of Na4SiO4, especially in applications such as solid-state batteries. The direct experimental characterization of these pathways is challenging due to the complexity of the material’s structure and the small scale of ion movements. Therefore, computational methods are often employed to simulate and analyze these diffusion pathways and activation energies. The lithium-ion migration path calculated in LiFePO4 using classical pair potentials was later exactly observed in high-temperature powder neutron diffraction and the maximum entropy method. The experiment visualized a curved one-dimensional chain for lithium motion with a Li-Li separation of 3.01 Å along the [010] direction [44]. This finding emphasizes the accuracy of our simulation methods in predicting ion migration pathways.
Using current simulation techniques, we constructed the Na-ion diffusion channels with atomic-level activation energies. Seven local Na hops were identified (A, B, C, D, E, F, and G) (see Figure 3). Hop A has a lower activation energy of 0.23 eV, indicating that it is a favorable and energetically accessible pathway for Na⁺ ion migration. Hop B has a moderate activation energy, suggesting it is less favorable compared to Hop A but still relatively accessible (see Table 3). Hop C, with the lowest activation energy among the hops listed, makes it the most energetically favorable for Na⁺ ion movement. Hop D has a moderate activation energy, making it less favorable than Hop A and C but more favorable than Hops D and F. This hop has a moderate activation energy, making it less favorable than Hop A and C but more favorable than Hops D and F. The Hop E pathway has the highest activation energy among the listed hops, indicating that it is the least favorable for Na⁺ ion migration. Hop G has a moderate activation energy, suggesting that it is less favorable than Hop A and C but more favorable than Hop D and F. The energy profile diagrams (see Figure 4) are essential for understanding the activation energies associated with different sodium-ion (Na⁺) diffusion pathways in Na4SiO4. These diagrams illustrate the energy barriers that Na⁺ ions must overcome to move through the material, providing insights into the ease or difficulty of diffusion.
We constructed and evaluated five promising long-range pathways by examining local Na⁺ hops and their associated activation energies (see Figure 4 and Table 4). The most favorable long-range path with the lowest overall activation energy (0.55 eV) is A↔C↔E↔A. In this long-range migration, Na+ ions migrate in the ac plane. An overall activation energy of 0.73 eV for three long-range pathways suggests the presence of mechanisms or reactions in materials or processes where the energy barrier for these pathways is also relatively low, facilitating easier transition or movement over longer distances. A long-range diffusion pathway with an overall migration energy of 1.02 eV indicates a significant energy barrier that must be overcome for atoms or ions to move through a material over extended distances. This high barrier can impact the material’s performance in various applications.

3.4. Solution of Dopants

Doping in battery materials is a strategic method to enhance various properties of the materials used in batteries, including their electrical conductivity, thermal stability, capacity, and cycle life. This process involves the intentional introduction of impurities (dopants) into the host material to improve its performance. When considering isovalent and aliovalent dopants in Na4SiO4, it is important to understand the impact these dopants have on the material’s structure and properties. Isovalent dopants have the same valence as the ions they replace, whereas aliovalent dopants have a different valence, which can lead to charge imbalances that must be compensated for within the crystal structure. Buckingham potentials used for dopants are provided in the electronic Supplementary Materials (ESM).

3.4.1. Monovalent Dopants

Monovalent dopants (M = Li, K, and Rb) in Na4SiO4 can alter its properties, including ionic conductivity and structural stability. Monovalent dopants have a single positive charge and can substitute for Na⁺ in the crystal lattice. The doping process is described by the following equation:
M 2 O + 2   N a N a X 2   M N a X + N a 2 O
Endoergic (or endothermic) solution enthalpies calculated for all dopants mean that the process absorbs energy from the surroundings. This implies that adding dopants to Na4SiO4 requires thermal energy, and the process is more thermodynamically favorable at higher temperatures. Among the dopants considered, K⁺ has the lowest solution enthalpy, meaning it requires the least amount of energy to be incorporated into Na4SiO4 (see Figure 5). This makes K⁺ the most thermodynamically favorable dopant. A dopant with a size closer to that of the host ion will cause less distortion in the lattice. K⁺, despite being larger, may fit into the Na4SiO4 lattice with less strain compared to other dopants, leading to a more stable structure. This can be advantageous for practical doping processes as it suggests easier incorporation and potentially better stability.

3.4.2. Trivalent Dopants

Incorporating trivalent dopants on the silicon site in Na4SiO4 can introduce various structural and electronic modifications that may enhance the material’s properties, such as ionic conductivity and stability. Substituting a Si4⁺ ion with a trivalent ion (M3⁺) introduces a negative charge in the lattice. This can be compensated for by creating oxygen vacancies or by incorporating additional sodium ions (Na⁺) to maintain charge neutrality. The creation of oxygen vacancies can enhance ionic conductivity by providing pathways for ion migration. Additionally, extra sodium ions introduced for charge compensation can increase the number of mobile ions according to the following equation. In this study, a diverse array of trivalent dopants (M = Al, Ga, Gd, In, Sc, and Y) were examined for incorporation at the Si site.
M 2 O 3 + 2   S i S i X + N a 2 O 2   M S i + 2   N a i + 2   S i O 2
Aluminum is the most promising trivalent dopant for improving Na4SiO4 due to its low solution energy and relatively small ionic radius (see Figure 6). The ionic radius of Al3⁺ is relatively close to that of Si4⁺. This similarity minimizes lattice distortions and strain when Al substitutes for Si. Minimal strain means the crystal structure remains stable and intact, making Al a favorable dopant. In previous simulation studies [45,46], Al has been reported to be the most promising dopant on the Si site in a variety of silicate materials. Gallium is also a promising dopant with the second lowest solution energy, indicating that it can also be incorporated into the Na4SiO4 lattice. Indium, despite having a larger ionic radius, has a solution energy slightly lower than Sc, but higher than Al and Ga, making it a moderately feasible dopant. Scandium, Yttrium, and Gadolinium, with higher solution energies and larger ionic radii, are less favorable for doping Na4SiO4. They introduce more significant lattice distortions and require more energy to incorporate, which may result in less effective enhancement of ionic conductivity.

3.4.3. Tetravalent Dopants

Tetravalent dopants (M = Ge, Sn, Ti, and Ce) on the Si site in Na4SiO4 can be used to modify and enhance the material’s properties while maintaining the charge balance, as they have the same +4 valence as silicon. The following reaction equation describes the following process:
M O 2 + S i S i X M S i X + S i O 2
Germanium has the lowest solution energy, indicating that it is the easiest dopant to incorporate into the Na4SiO4 lattice (see Figure 7). The ionic radius of Ge4⁺ (0.53 Å) is close to that of Si4⁺ (0.40 Å), ensuring minimal lattice distortion. This makes Ge4⁺ an excellent choice for doping, enhancing ionic conductivity, and maintaining structural integrity. Tin has a moderate solution energy, suggesting that it is relatively easy to incorporate but not as easy as Ge4⁺. Sn4⁺ can be beneficial for certain applications, but its larger ionic radius (0.69 Å) compared to Si4⁺ could introduce some lattice distortions. Titanium has a higher solution energy, making it more difficult to incorporate into the lattice. Ti4⁺ has an ionic radius of 0.61 Å, which is larger than Si4⁺, potentially causing lattice distortions. However, Ti4⁺ can enhance mechanical properties. Cerium has the highest solution energy, indicating the greatest difficulty in incorporating it into the lattice. Ce4⁺ has a larger ionic radius, which can lead to significant lattice distortions. It may provide unique properties but is less favorable due to its high solution energy.

4. Conclusions

In this study, we used atomistic simulation techniques to examine the defect, diffusion, and doping properties of a promising battery material, Na4SiO4. The Na Frenkel defect was found to be the most energetically favorable defect process facilitating vacancy-assisted Na-ion migration. Three-dimensional Na⁺ ion diffusion is characterized by an activation energy of 0.55 eV. K and Ge are favorable dopants for the Na and Si sites in Na4SiO4 respectively. The doping of the Si site in Na4SiO4 with trivalent Al3⁺ ions is favorable, generating additional Na⁺ ions in the material. Experimental validation of these findings will be crucial in confirming their practical implications and optimizing Na4SiO4 for high-performance battery applications.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/cryst14080718/s1. Table S1: Buckingham potential parameters used for dopants in the atomistic simulations of Na4SiO4.

Author Contributions

Conceptualization, M.S.S.; methodology, M.S.S.; validation, P.A., P.I. and N.K.; formal analysis, M.S.S.; investigation, M.S.S.; writing—original draft preparation, N.K.; writing—review and editing, P.A.; supervision, P.A.; project administration. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

All data are provided either in the main article or in the electronic Supplementary Materials.

Acknowledgments

We acknowledge computational facilities provided by the Department of Chemistry, University of Jaffna, Sri Lanka, and Imperial College London, UK.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. 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]
  2. Panda, N.; Cueva-Sola, A.B.; Dzulqornain, A.M.; Thenepalli, T.; Lee, J.-Y.; Yoon, H.-S.; Jyothi, R.K. Review on lithium-ion battery recycling: Challenges and possibilities. Geosystem Eng. 2023, 26, 101–118. [Google Scholar] [CrossRef]
  3. 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]
  4. Feng, X.; Ren, D.; Ouyang, M. Safety of lithium battery materials chemistry. J. Mater. Chem. A 2023, 11, 25236–25246. [Google Scholar] [CrossRef]
  5. Grey, C.P.; Hall, D.S. Prospects for lithium-ion batteries and beyond—A 2030 vision. Nat. Commun. 2020, 11, 6279. [Google Scholar] [CrossRef] [PubMed]
  6. Halder, P.; Bhattacharya, S. Battery Applications. In Lithium Ion Glassy Electrolytes: Properties, Fundamentals, and Applications; Bhattacharya, S., Bhattacharya, K., Eds.; Springer Nature: Singapore, 2022; pp. 159–173. [Google Scholar]
  7. Frith, J.T.; Lacey, M.J.; Ulissi, U. A non-academic perspective on the future of lithium-based batteries. Nat. Commun. 2023, 14, 420. [Google Scholar] [CrossRef] [PubMed]
  8. Ralls, A.M.; Leong, K.; Clayton, J.; Fuelling, P.; Mercer, C.; Navarro, V.; Menezes, P.L. The Role of Lithium-Ion Batteries in the Growing Trend of Electric Vehicles. Materials 2023, 16, 6063. [Google Scholar] [CrossRef] [PubMed]
  9. Kennedy, B.; Patterson, D.; Camilleri, S. Use of lithium-ion batteries in electric vehicles. J. Power Sources 2000, 90, 156–162. [Google Scholar] [CrossRef]
  10. Chen, W.; Liang, J.; Yang, Z.; Li, G. A Review of Lithium-Ion Battery for Electric Vehicle Applications and Beyond. Energy Procedia 2019, 158, 4363–4368. [Google Scholar] [CrossRef]
  11. Wu, H.; Hao, J.; Jiang, Y.; Jiao, Y.; Liu, J.; Xu, X.; Davey, K.; Wang, C.; Qiao, S.-Z. Alkaline-based aqueous sodium-ion batteries for large-scale energy storage. Nat. Commun. 2024, 15, 575. [Google Scholar] [CrossRef]
  12. Gupta, P.; Pushpakanth, S.; Haider, M.A.; Basu, S. Understanding the Design of Cathode Materials for Na-Ion Batteries. ACS Omega 2022, 7, 5605–5614. [Google Scholar] [CrossRef] [PubMed]
  13. Skundin, A.M.; Kulova, T.L.; Yaroslavtsev, A.B. Sodium-Ion Batteries (a Review). Russ. J. Electrochem. 2018, 54, 113–152. [Google Scholar] [CrossRef]
  14. Ferraro, M.; Tumminia, G. Techno-economics Analysis on Sodium-Ion Batteries: Overview and Prospective. In Emerging Battery Technologies to Boost the Clean Energy Transition: Cost, Sustainability, and Performance Analysis; Passerini, S., Barelli, L., Baumann, M., Peters, J., Weil, M., Eds.; Springer International Publishing: Berlin/Heidelberg, Germany, 2024; pp. 259–266. [Google Scholar]
  15. Que, L.; Yu, F.; Wu, J.; Lan, Z.; Feng, Y.; Zhao, R.; Sun, Z.; Yang, Z.; Luo, H.; Chao, D. Unveil the origin of voltage oscillation for sodium-ion batteries operating at −40 °C. Proc. Natl. Acad. Sci. USA 2024, 121, e2311075121. [Google Scholar] [CrossRef] [PubMed]
  16. Mamoor, M.; Li, Y.; Wang, L.; Jing, Z.; Wang, B.; Qu, G.; Kong, L.; Li, Y.; Guo, Z.; Xu, L. Recent progress on advanced high energy electrode materials for sodium ion batteries. Green Energy Resour. 2023, 1, 100033. [Google Scholar] [CrossRef]
  17. Singh, B.; Wang, Z.; Park, S.; Gautam, G.S.; Chotard, J.-N.; Croguennec, L.; Carlier, D.; Cheetham, A.K.; Masquelier, C.; Canepa, P. A chemical map of NaSICON electrode materials for sodium-ion batteries. J. Mater. Chem. A 2021, 9, 281–292. [Google Scholar] [CrossRef]
  18. Jayamkondan, Y.; Penki, T.R.; Nayak, P.K. Recent advances and challenges in the development of advanced positive electrode materials for sustainable Na-ion batteries. Mater. Today Energy 2023, 36, 101360. [Google Scholar] [CrossRef]
  19. Bianchini, F.; Fjellvåg, H.; Vajeeston, P. First-principles study of the structural stability and electrochemical properties of Na2MSiO4 (M = Mn, Fe, Co and Ni) polymorphs. Phys. Chem. Chem. Phys. 2017, 19, 14462–14470. [Google Scholar] [CrossRef] [PubMed]
  20. Zhu, L.; Li, L.; Wen, J.; Zeng, Y.-R. Structural stability and ionic transport property of NaMPO4 (M = V, Cr, Mn, Fe, Co, Ni) as cathode material for Na-ion batteries. J. Power Sources 2019, 438, 227016. [Google Scholar] [CrossRef]
  21. Strauss, F.; Rousse, G.; Sougrati, M.T.; Dalla Corte, D.A.; Courty, M.; Dominko, R.; Tarascon, J.-M. Synthesis, Structure, and Electrochemical Properties of Na3MB5O10 (M = Fe, Co) Containing M2+ in Tetrahedral Coordination. Inorg. Chem. 2016, 55, 12775–12782. [Google Scholar] [CrossRef]
  22. Sivakumaran, A.; Samson, A.J.; Bristi, A.A.; Surendran, V.; Butler, S.; Reid, S.; Thangadurai, V. High ionic conducting rare-earth silicate electrolytes for sodium metal batteries. J. Mater. Chem. A 2023, 11, 15792–15801. [Google Scholar] [CrossRef]
  23. Baur, W.H.; Halwax, E.; Völlenkle, H. Comparison of the crystal structures of sodium orthosilicate, Na4SiO4, and sodium orthogermanate, Na4GeO4. Monatshefte Chem./Chem. Mon. 1986, 117, 793–797. [Google Scholar] [CrossRef]
  24. Ling, C.; Luo, X.; Wang, Z.; Tang, Z. Experimental and DFT study of Na4SiO4 doped with oxysalts for high-temperature CO2 capture. Chem. Eng. J. 2024, 495, 153331. [Google Scholar] [CrossRef]
  25. Ling, C.; Wang, Z.; Tang, Z. Performance improvement of Na4SiO4 doped with Li2CO3-K2CO3 for high-temperature CO2 capture and thermochemical energy storage. Chem. Eng. J. 2023, 476, 146921. [Google Scholar] [CrossRef]
  26. Plascencia-Hernández, F.; Araiza, D.G.; Pfeiffer, H. Effect of Sodium Ortho and Pyrosilicates (Na4SiO4–Na6Si2O7) Mixture during the CO2 Chemical Capture Performance. Ind. Eng. Chem. Res. 2022, 61, 11012–11024. [Google Scholar] [CrossRef]
  27. Wang, Z.; Sun, C.; Xu, Q.; Zou, X.; Cheng, H.; Lu, X. In Situ XRD, Raman Characterization, and Kinetic Study of CO2 Capture by Alkali Carbonate-Doped Na4SiO4. Separations 2022, 9, 428. [Google Scholar] [CrossRef]
  28. Liu, J.; Wang, Z.; Wang, Z.; Song, J.; Li, G.; Xu, Q.; You, J.; Cheng, H.; Lu, X. Alkali carbonates promote CO2 capture by sodium orthosilicate. Phys. Chem. Chem. Phys. 2019, 21, 13135–13143. [Google Scholar] [CrossRef] [PubMed]
  29. Aparicio, P.A.; Dawson, J.A.; Islam, M.S.; de Leeuw, N.H. Computational Study of NaVOPO4 Polymorphs as Cathode Materials for Na-Ion Batteries: Diffusion, Electronic Properties, and Cation-Doping Behavior. J. Phys. Chem. C 2018, 122, 25829–25836. [Google Scholar] [CrossRef]
  30. Kuganathan, N.; Kordatos, A.; Anurakavan, S.; Iyngaran, P.; Chroneos, A. Li3SbO4 lithium-ion battery material: Defects, lithium ion diffusion and tetravalent dopants. Mater. Chem. Phys. 2019, 225, 34–41. [Google Scholar] [CrossRef]
  31. Kumar Prajapati, A.; Bhatnagar, A. A review on anode materials for lithium/sodium-ion batteries. J. Energy Chem. 2023, 83, 509–540. [Google Scholar] [CrossRef]
  32. Goldmann, B.A.; Clarke, M.J.; Dawson, J.A.; Islam, M.S. Atomic-scale investigation of cation doping and defect clustering in the anti-perovskite Na3OCl sodium-ion conductor. J. Mater. Chem. A 2022, 10, 2249–2255. [Google Scholar] [CrossRef]
  33. Ahiavi, E.; Dawson, J.A.; Kudu, U.; Courty, M.; Islam, M.S.; Clemens, O.; Masquelier, C.; Famprikis, T. Mechanochemical synthesis and ion transport properties of Na3OX (X = Cl, Br, I and BH4) antiperovskite solid electrolytes. J. Power Sources 2020, 471, 228489. [Google Scholar] [CrossRef]
  34. Islam, M.S.; Fisher, C.A.J. Lithium and sodium battery cathode materials: Computational insights into voltage, diffusion and nanostructural properties. Chem. Soc. Rev. 2014, 43, 185–204. [Google Scholar] [CrossRef]
  35. 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]
  36. Gale, J.D.; Rohl, A.L. The General Utility Lattice Program (GULP). Mol. Simul. 2003, 29, 291–341. [Google Scholar] [CrossRef]
  37. 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]
  38. 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] [PubMed]
  39. Kuganathan, N.; Ganeshalingam, S.; Chroneos, A. Defects, Diffusion, and Dopants in Li2Ti6O13: Atomistic Simulation Study. Materials 2019, 12, 2851. [Google Scholar] [CrossRef]
  40. Kuganathan, N.; Solovjov, A.L.; Vovk, R.V.; Chroneos, A. Defects, diffusion and dopants in Li8SnO6. Heliyon 2021, 7, e07460. [Google Scholar] [CrossRef]
  41. Heath, J.; Chen, H.; Islam, M.S. MgFeSiO4 as a potential cathode material for magnesium batteries: Ion diffusion rates and voltage trends. J. Mater. Chem. A 2017, 5, 13161–13167. [Google Scholar] [CrossRef]
  42. Pedone, A.; Gambuzzi, E.; Malavasi, G.; Menziani, M.C. First-principles simulations of the 27Al and 17O solid-state NMR spectra of the CaAl2Si3O10 glass. Theor. Chem. Acc. 2012, 131, 1147. [Google Scholar] [CrossRef]
  43. 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]
  44. Fisher, C.A.J.; Hart Prieto, V.M.; 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]
  45. Islam, M.S.; Dominko, R.; Masquelier, C.; Sirisopanaporn, C.; Armstrong, A.R.; Bruce, P.G. Silicate cathodes for lithium batteries: Alternatives to phosphates? J. Mater. Chem. 2011, 21, 9811–9818. [Google Scholar] [CrossRef]
  46. Jones, A.; Slater, P.R.; Islam, M.S. Local Defect Structures and Ion Transport Mechanisms in the Oxygen-Excess Apatite La9.67(SiO4)6O2.5. Chem. Mater. 2008, 20, 5055–5060. [Google Scholar] [CrossRef]
Figure 1. Crystal structure of Na4SiO4.
Figure 1. Crystal structure of Na4SiO4.
Crystals 14 00718 g001
Figure 2. Defect energies of different defect processes in Na4SiO4.
Figure 2. Defect energies of different defect processes in Na4SiO4.
Crystals 14 00718 g002
Figure 3. Seven different Na local hops considered for the long-range migration pathways.
Figure 3. Seven different Na local hops considered for the long-range migration pathways.
Crystals 14 00718 g003
Figure 4. Energy profile diagrams calculated for different local Na hops.
Figure 4. Energy profile diagrams calculated for different local Na hops.
Crystals 14 00718 g004
Figure 5. Solution energies calculated for M2O (M = Li, K, Rb).
Figure 5. Solution energies calculated for M2O (M = Li, K, Rb).
Crystals 14 00718 g005
Figure 6. Solution energies calculated for M2O3 (M = Al, Ga, In, Sc, Y, and Gd).
Figure 6. Solution energies calculated for M2O3 (M = Al, Ga, In, Sc, Y, and Gd).
Crystals 14 00718 g006
Figure 7. Solution energies calculated for MO2 (M = Ge, Sn, Ti, and Ce).
Figure 7. Solution energies calculated for MO2 (M = Ge, Sn, Ti, and Ce).
Crystals 14 00718 g007
Table 1. Buckingham potential parameters used in the atomistic simulations of Na4SiO4 [40,41,42]. 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 atomistic simulations of Na4SiO4 [40,41,42]. 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/eVρC/eV·Å6Y/eK/eV·Å–2
Na+–O2−1497.8305980.2874830.001.0099,999
Si4+–O2−1283.910.3205210.664.0099,999
O2—O2−22,764.300.149027.88−2.8661.50
Table 2. Calculated and experimental lattice parameters of Na4SiO4.
Table 2. Calculated and experimental lattice parameters of Na4SiO4.
ParameterCalculatedExperiment [23]|Δ|%
a (Å)5.565.580.33
b (Å)5.625.580.77
c (Å)8.398.390.07
α (°)81.3480.920.52
β (°)71.8771.840.04
γ (°)67.0267.440.62
Table 3. Local Na hops and their activation energies.
Table 3. Local Na hops and their activation energies.
Na-Na HopDistance (Å)Activation Energy (eV)
A3.14 Å0.23
B3.62Å0.65
C3.39 Å0.03
D3.89Å0.73
E3.29 Å0.55
F4.07Å1.02
G2.91 Å0.62
Table 4. Long-range Na ion diffusion pathways and their overall activation energies.
Table 4. Long-range Na ion diffusion pathways and their overall activation energies.
Long-Range PathwayHop Activation Energies (eV)Overall Activation Energy (eV)
A↔C↔E↔A0.23 ↔0.03↔0.55↔0.230.55
D↔E↔G↔D↔G0.73 ↔0.55↔0.62↔0.73↔0.620.73
A↔C↔E↔D↔G0.23 ↔0.03↔0.55↔0.73↔0.620.73
A↔B↔E↔G↔D↔G0.23 ↔0.65↔0.55↔0.62↔0.73↔0.620.73
A↔C↔E↔F↔A0.23 ↔0.03↔0.55↔1.02↔0.231.02
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

Shanika, M.S.; Abiman, P.; Iyngaran, P.; Kuganathan, N. Atomistic Simulation Studies of Na4SiO4. Crystals 2024, 14, 718. https://doi.org/10.3390/cryst14080718

AMA Style

Shanika MS, Abiman P, Iyngaran P, Kuganathan N. Atomistic Simulation Studies of Na4SiO4. Crystals. 2024; 14(8):718. https://doi.org/10.3390/cryst14080718

Chicago/Turabian Style

Shanika, Mallikage Shalani, Poobalasingam Abiman, Poobalasuntharam Iyngaran, and Navaratnarajah Kuganathan. 2024. "Atomistic Simulation Studies of Na4SiO4" Crystals 14, no. 8: 718. https://doi.org/10.3390/cryst14080718

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