Next Article in Journal
Study of Endocrine-Disrupting Chemicals in Infant Formulas and Baby Bottles: Data from the European LIFE-MILCH PROJECT
Previous Article in Journal
NMR Relaxation to Probe Zeolites: Mobility of Adsorbed Molecules, Surface Acidity, Pore Size Distribution and Connectivity
Previous Article in Special Issue
Efficient Removal of Cationic Dye by Biomimetic Amorphous Calcium Carbonate: Behavior and Mechanisms
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

First-Principles Study of 3R-MoS2 for High-Capacity and Stable Aluminum Ion Batteries Cathode Material

1
School of Physics and Electronic Engineering, Xinxiang University, Xinxiang 453003, China
2
School of Mechanical Engineering, Chengdu University, Chengdu 610106, China
3
Henan Province Engineering Research Center of New Energy Storage System, Xinxiang University, Xinxiang 453003, China
4
School of Physical Science and Technology, Lanzhou University, Lanzhou 730000, China
*
Author to whom correspondence should be addressed.
Molecules 2024, 29(22), 5433; https://doi.org/10.3390/molecules29225433
Submission received: 11 October 2024 / Revised: 9 November 2024 / Accepted: 16 November 2024 / Published: 18 November 2024
(This article belongs to the Special Issue Functional Nanomaterials for Energy and Environmental Sustainability)

Abstract

:
Currently, exploring high-capacity, stable cathode materials remains a major challenge for rechargeable Aluminum-ion batteries (AIBs). As an intercalator for rechargeable AIBs, Al3+ produces three times the capacity of AlCl4 when the same number of anions is inserted. However, the cathode material capable of producing Al3+ intercalation is not a graphite material with AlCl4 intercalation but a transition metal sulfide material with polar bonding. In this paper, the insertion mechanism of Al3+ in 3R-MoS2 is investigated using first-principles calculations. It is found that Al3+ tends to insert into different interlayer positions at the same time rather than occupying one layer before inserting into another, which is different from the insertion mechanism of AlCl4 in graphite. Ab initio, molecular dynamics calculations revealed that Al3+ was able to stabilize the insertion of 3R-MoS2. Diffusion barriers indicate that Al3+ preferentially migrates to nearby stabilization sites in diffusion pathway studies. According to the calculation, the theoretical maximum specific capacity of Al3+ intercalated 3R-MoS2 reached 502.30 mAg h−1, and the average voltage of the intercalation was in the range of 0.75–0.96 V. Therefore, 3R-MoS2 is a very promising cathode material for AIBs.

1. Introduction

Energy storage batteries have always been the focus of attention. Currently, metal-ion batteries have attracted significant attention from researchers due to their small size and high efficiency. Lithium-ion batteries (LIBs) are a widely utilized energy storage solution due to their high energy density, charge/discharge voltage, and capacity [1,2]. However, in comparison to other metal resources, lithium metal reserves are relatively limited (0.0017 wt%), and the cost of LIBs has been on an upward trajectory. Furthermore, safety concerns, such as the formation of lithium dendrites, are becoming increasingly prominent [3,4,5,6,7]. In this context, Aluminum-ion batteries (AIBs) have emerged as a highly promising option due to their abundance of energy storage, high energy density, and extensive reserves [8]. However, AIBs still lack cathode materials with high capacity and stability [9,10].
In 2015, Dai et al. developed ultra-fast charging AIBs comprising an anode of metallic aluminum and a cathode of three-dimensional foam graphite, a material that exhibits an ultra-fast charging rate but a low capacity [11]. This was due to the intercalation and deintercalation of AlCl4 in the electrode material, which did not utilize the properties of Al3+. Furthermore, the large size of the chloroaluminum anion causes volume expansion [12]. Al3+ carries more charge than AlCl4, which means that when the same amount of AlCl4 is inserted, Al3+ can produce more capacity. Currently, transition metal oxides with polar bonds are considered potential cathode materials for Al3+ inserted [13]. In 2011, Jayaprakash et al. reported for the first time that V2O5 nanowires could serve as cathode materials for Al3+ intercalation. The cell obtained a capacity retention of 273 mAh g−1 after 20 cycles at a current density of 125 mA g−1. However, the high charge density on the Al3+ surface produces strong Coulombic ion−lattice interactions that hinder the diffusion of Al3+ in the lattice, limiting the rate performance and cycling stability of the AIBs [13,14]. To reduce this strong Coulombic ion–lattice interaction, transition metal sulfides (TMDs) with weak Coulombic interactions have been used for extensive studies [15,16]. In 2018, Li et al. first prepared a rechargeable aluminum ion battery consisting of MoS2 microsphere cathode, aluminum anode, and ionic liquid electrolyte, demonstrating the feasibility of MoS2 as a cathode for AIBs [17]. In 2019, Tu et al. synthesized hierarchical flower-like MoS2 microspheres as AIBs cathodes by a simple hydrothermal method and showed that the structure of MoS2 was stabilized during Al3+ intercalation/deintercalation by density functional theory calculations [18]. All the above studies are based on 2H-MoS2. However, 3R-MoS2, which is also a semiconductor phase with a stable thermodynamic ABC stacking mode (unlike 2H-MoS2 AB stacking), lacks studies specifically on electrochemical behavior [19,20].
This study systematically investigates the structural stability, electronic properties, theoretical capacity, and average voltage of Al3+ intercalation/deintercalation 3R-MoS2 electrodes using first-principles calculations. Furthermore, the thermal stability of 3R-MoS2 is examined using the Ab Initio Molecular Dynamics (AIMD) method, and the diffusion path of Al3+ in the 3R-MoS2 layer is investigated in detail. Our findings were benchmarked against the most recent experimental data on AIBs. Furthermore, we conducted a detailed investigation of the Al3+ inserted MoS2 system to elucidate the cation insertion mechanism in MoS2, with the objective of facilitating the design of superior electrodes in the future.

2. Results and Discussions

2.1. Single Al3+ Inserted in 3R-MoS2

A comprehensive examination of all potential inserted locations for Al3+ within MoS2 has been conducted. Two preferred insertion positions of Al3+ in 3R-MoS2 were envisioned based on the study of potassium ion intercalation at the 2H-MoS2 site (Figure 1): positions A and B (named PA and PB) [21]. At PA, the Al3+ is situated within an octahedron comprising six S atoms. It occupies a bridging position between two nonbonding S atoms, thereby forming six Al–S bonds. At PB, the Al3+ is situated within a tetrahedron of four S atoms, and Al3+ also occupies a bridging position between two non−bonded S atoms, forming four Al–S bonds. According to binding energy studies, these two positions are very favorable, and the two energies are very close. (Table S1). The binding energy for PA is less than that of PB, indicating that PA is the more stable site for Al3+ intercalation (consistent with the stabilization site in 2H-MoS2 [18]). The subsequent section of this study focuses on the insertion of an Al3+ at PA. The energies associated with the intercalation of an Al3+ into the top, middle, and bottom layers of PA exhibit minimal variation. This indicates that the initial insertion of the Al3+ is random with respect to the layer in question.
The thermal stability of Al3+ inserted into the 3R-MoS2 structure was investigated with AIMD (Figure S1). The most stable PA was selected for the simulation. Initially, the structure was subjected to a heating process, reaching a temperature of 300 K over a time interval of 1 fs to 5 ps. It was observed that the ABC stacking of 3R-MoS2 remained unaltered from the snapshot taken at 5 ps. The structure remained intact, and the embedded Al3+ exhibited minimal displacement within the PA. Furthermore, simulations were conducted using the NVT system at temperatures of 400 K, 500 K, and 600 K with a time step of 1 fs and a time step of 5 ps. It was observed that 3R-MoS2 also maintains the ABC stacking at temperatures between 300 K and 600 K. There is no significant displacement at the PA, and the changes in bond lengths and bond angles are not significant. The results demonstrate that intercalation in the PA is stable in AIMD simulations conducted at temperatures between 300 K and 600 K. Additionally, the Al3+ can diffuse rapidly into 3R-MoS2 due to the unaltered chemical bonding nature of the latter.

2.2. More Al3+ Inserted in 3R-MoS2

As shown in Figure 2 and Figure S2, we simulated three different interpolation stages (stages 1–3), each using four different concentrations. 3, 6, 9 and 12 Al3+ are inserted in MoS2 for stage-1, 2, 4, 6 and 8 Al3+ are inserted in MoS2 for stage-2. 1, 2, 3 and 4 Al3+ are inserted in MoS2 for stage-3. A 3 × 3 × 2 supercell consisting of 24 S atoms and 12 Mo atoms was constructed for the calculations of stage-1, stage-2, and stage-3. It should be noted that the system was modeled using a pure 3R-MoS2 ABC-stacked structure comprising three MoS2 layers with an interlayer spacing of 2.80 Å. It is, therefore, believed that the MoS2 model is more suitable for simulating real experimental observations in ultra−fast AIBs.
It has been found that good cathode materials for AIBs require suitable interlayer spacing as well as reasonable Al binding strength. Therefore, MoS2 with these two characteristics happens to be a potential cathode for AIBs. We first investigated the structural deformation in MoS2 due to Al3+ inserted. There are three interlayer distances in the unit cell named Ind1, Ind2, and Ind3, shown in Figure S3. All distances are projected onto the C-axis and listed in Table 1. The layer spacing in pure MoS2 (2.80 Å), which is very close to the experiment data 2.98 Å and corresponds to the (006) diffraction plane of MoS2 (JCPDS No. 17–0744), is larger than the covalent radius of Al (1.18 Å) [22]. The results show that the interlayer spacing expands with the insertion of Al3+ at first for all the stages. All the expansion layers spacing are about 3.2 Å. This is due to the electrostatic interaction between Al and Mo ions. And the distance shrinks 0.1–0.5 Å for the layer without Al3+ inserted. After the interlayer spacing expands, further intercalation becomes smooth. Thus, with more Al3+ continuing to be embedded, the change in MoS2 layer spacing becomes smaller. It can also be considered that the insertion of Al3+ will not cause a large change in layer spacing, which verified that the structure of 3R-MoS2 is stable. The bonding of the Al and S atoms also prevents further expansion of the layer spacing. The increased interlayer spacing favors easier diffusion of Al3+ into MoS2. However, when 9 Al3+ was inserted in MoS2 in stage-1, and 6 Al3+ was inserted in MoS2 in stage-2, the lattice changed much (Figure S2), and the interlayer spacing varied irregularly. For example, when 9 Al3+ is inserted in MoS2 in stage-1, the Ind1 and Ind2 become smaller. This implies that the electrostatic interaction force is not uniform. However, the calculated expanded layer spacing for 3R-MoS2 (1.1 Å) is smaller than the one of 2H-MoS2 (1.9 Å), which indicates that the structure stability of 3R-MoS2 is better than the one of 2H-MoS2 [17,23].

2.3. Staging Mechanism

To further analyze the staging mechanism, the relative stabilities of stage-1, stage-2, and stage-3 with the same Al3+ concentration were compared. Divided into four sections, each inserted with the same concentration of Al3+, the optimized structures of stages 1, 2, and 3 are shown in Figure S4 and expressed in terms of relative energetics. The results show that stage-1 is more stable than stage-2 when the same concentration of 6 Al3+ is inserted, and stage-2 is again more stable than stage-3 when Al3+ is reduced to 4, and the same is true for subsequent reductions to 3 and 2. Thus, we believe that Al3+ prefers to insert into the galleries simultaneously rather than covering one gallery and then taking over the others. This is different from the mechanism of AlCl4 insertion in graphene [24]. This phenomenon can be explained by the ion−ion Coulomb attraction. As Al ion and S ion bond to each other, it is believed that Al3+ first binds to S, and with the Al3+ increasing, the repulsion between Al and Mo atoms increases. Thus, Al3+ will not cover one layer and then take over the other. That will increase the repulsive force between Al and Mo atoms. As the concentration of Al3+ increases, the repulsive force also rises, which may lead to lattice distortion, such as the formation of 6 Al3+. Nevertheless, if Al3+ is distributed uniformly throughout all layers, it will result in a reduction in lattice distortion, such as the presence of 6 Al3+ in stage-1. It is, therefore, postulated that the intercalation mechanism follows stage-1.

2.4. Binding Energy

To further assess the stability of the intercalated compounds, the binding energies calculated for all stages are listed in Table 1, and Figure S5 shows the variation of binding energy as a function of the weight percentage (wt%) of Al3+ intercalation in MoS2 for all three stages. The calculated binding energies are observed to be negative in all cases, indicating that Al3+ is readily embedded in MoS2 within the context of this study. This phenomenon can be attributed to the smaller atomic size of Al3+ in comparison to the spacing of the MoS2 layers, which facilitates the insertion process. Moreover, the interaction between the intercalated Al3+ and the host MoS2 layer is sufficient to overcome the van der Waals forces between the MoS2 layers. In addition, the insertion of Al3+ in the already expanded MoS2 main channel is facilitated by higher concentrations. This is due to an increase in the average interlayer distance, which results in a reduction in van der Waals forces between the MoS2 layers. It can be demonstrated that the Al3+ binding energy in the bulk MoS2 decreases with the amount of Al3+ in the same stage. Furthermore, the lower stages (stage-1) are more stable than the higher stages (stage-2, stage-3) for a given concentration of Al3+.

2.5. Electronic Properties

Conductive properties are very important for battery materials. The total density of states (TDOSs) and partial density of states (PDOSs) are calculated for Al3+ intercalated in MoS2 of stage-1, as shown in Figure 3. Only DOSs near the Fermi level are meaningful. Thus, the DOSs between −10 and 15 eV are shown in Figure 3. Furthermore, the DOSs at the Fermi level can effectively reflect the conductivity of the material [25]. It can be seen that with no Al3+ intercalating, 3R-MoS2 has poor electrical conductivity. Because there are fewer electrons at the Fermi level. And the electrons are mainly from Mo 4d and S 3p, which can be found in Figure S6a. The parts from −10 eV to 0 eV mainly are Mo 4d and S 3p hybridization, forming the S–Mo bond. With Al3+ intercalating, the DOSs at the Fermi level increase, which means the electronic conductivity increases. The electronic conductivity increases are mainly from Al 3p, Mo 4d, and S 3p orbitals. It is supposed that when Al3+ intercalates, the electrons transfer from Al 3p to S 3p. Correspondingly, the charge transfer of the Mo 4d is reduced, which can be verified from the Hirshfeld charge analysis. Thus, during charge/discharge, Al3+ increases the electronic conductivity of the materials.
Hirshfeld charge analysis is performed to understand the charge transfer more intuitively. From Figure S7, it was found that with the insertion of Al3+, the positive charge of Al increases and the charge of Mo decreases, whereas the Al metal is conductive and tends to lose its outermost electrons, which also leads to the transfer of electrons from the outermost layer of Al to S, resulting in the formation of Al–S bonds. And the electron transfer of Mo is weakening. Thus, the Mo–S bond is weakening. The electrons mainly transfer from Al 3p to S 3p orbitals. Furthermore, part electrons of Al 3s jump to Al 3p. The same goes for S atoms. Thus, both s and p orbitals of Al and S are hybridized and participate in bonding, which can be verified from Figure 3 and Figure S6b. The electrons transfer for Mo are mainly from Mo 5s to S 3p orbitals. And part electrons of Mo 5s jump to Mo 4d. Thus, there are fewer electrons left in the Mo 5s orbital. Only Mo 4d bond with S 3p, which agreed well with the analysis of the DOSs above. The sudden increase in Hirshfeld’s charge is due to lattice distortion for 9 Al3+ inserted in MoS2.

2.6. Diffusion of Al3+ in 3R-MoS2

The charging/discharging rate of batteries is related to the diffusion of the intercalated species through the electrodes. Therefore, to study the Al3+ diffusion mechanism in the 3R-MoS2 electrode, we have calculated barriers to finding energetically favorable diffusion paths. The PA site is the most stable. Hence, in this work, three favorable diffusion paths have been studied from a PA site to a neighboring A site. The first diffusion path, path 1, is from site PA to its nearest PA site, passing through a PB1 site (PA-PB1-PA); the second is path 2 involves the Al3+ diffusion from site A to another A site via site PB2 (PA-PB2-PA) and the third path, path 3, from site A to another nearest site A via site PB3 (PA-PB3-PA). All the diffusion paths and diffusion barriers are shown in Figure 4. From Figure 4, it is found that path 2 has the biggest barrier energy, as the diffusion path is the longest. Moreover, the barrier energies of path 1 and path 3 are closer and smaller than the ones of path 2. This can indicate that aluminum ions preferentially migrate to nearby stable positions. Moreover, the attraction of the S atom to Al3+ increased resistance and led to a higher barrier energy.

2.7. Electrochemical Properties of Al3+ Intercalated 3R-MoS2

The open circuit voltage (OCV) is a valuable measure of a battery’s performance. The open-circuit voltage (OCV) is evaluated by the equation below [26]:
V = ( E x 2 E x 1 ( x 1 x 2 ) E A l ) / 3 e ( x 1 x 2 )
where E x 1 and E x 2 are the total energies of the AlxMoS2 at two adjacent stable systems of concentration x1 and x2, respectively.
Figure 5 represents the variation of average voltage with the weight percentage (wt%) of Al3+ for all three stages (stages 1–3), and the calculated values for different stoichiometries are listed in Table 1. The OCV varies in a range of 0.77 V to 0.96 V for stage-1, 0.81 V to 0.94 V for stage-2, and 0.75 V to 0.92 V for stage-3. Since Al3+ are inserted into each layer at the same time in stage-1, the Coulomb and van der Waals forces of each layer need to be overcome, but stage-2 and stage-3 do not, which results in a higher voltage plateau in stage-1 relative to stage-2 and stage-3 [17,27,28].
The theoretical gravimetric capacity has been calculated using the following equation [29]:
C = c z F / M M o S 2
where c is the number of inserted Al3+, z is the number of chemical valences of Al3+, F is the Faraday constant (26,801 mA h mol−1) and M M o S 2 is the molar mass of the MoS2 cell. The calculated gravimetric capacities for stage-1, stage-2, and stage-3 are 502.30, 334.87, and 167.43 mAh g−1, listed in Table 2. From Table 2, it is obvious that the theoretical capacity of 3R-MoS2 stage-1 is higher than the experimental data of 2H-MoS2. Furthermore, the discharge plateau is observed to be higher than that observed in the other experiments. Therefore, based on our research and analysis, 3R-MoS2 is a potential cathode material for AIBs. The material’s stable layered structure enables it to maintain a long cycle life, high platform, and high capacity.

3. Computational Methods

The current calculations are based on Density Functional Theory (DFT) in the Cambridge Series of Total Energy Packets (CASTEP) plane wave code and on DS-PAW (Device Studio-Projected Augmented Wave) [33,34,35,36,37]. The Predew–Burke–Ernzerhof (PBE) generalized gradient approximation (GGA) was used in the description of the exchange–correlation potential. Norm-conserving pseudopotentials were used to describe the interaction of ionic core and valence electrons. Valence states were considered in this study corresponding to Al 3s2 3p1 Mo 4d5 5s1 and S 3s2 3p4. After convergence tests (Figure S8), the cutoff energy for plane waves is eventually identified as 1820 eV. The d orbitals in the Mo ions have been localized using the Hubbard corrected GGA (GGA + U) method (U value of 4 eV). All structures are optimized by fully relaxing the atomic and lattice positions until the Hermann−Feynman force for all atoms is less than 0.01 eV Å−1. The numbers of K-points were set to 5 × 5 × 1. All systems are fully optimized, and the convergence criterion for the total energy is set to 5 × 10−3 eV. Van der Waals force interactions play an important role in the calculation of layered structures. Therefore, we have corrected the van der Waals force interactions using the DFT-3 method, which increases the potential energy and interatomic forces.
In this paper, a periodic structure was adopted. The unit cell of 3R-MoS2 is of the R3m space group and exhibits a hexagonal structure (Figure S3). The 2 × 2 × 2 supercell was employed in the calculations, which included 12 Mo atoms and 36 S atoms. Following the completion of geometry optimization, the resulting geometry constants are presented in Table S1. As evidenced by Table S2, the findings align with those of other experimental and theoretical studies [38,39,40].
To calculate the electronic structure, a 4 × 4 × 4 K-point sampling was employed within the Brillouin region. Ab initio molecular dynamics simulations (AIMD) were conducted using a regular ensemble (NVT) with a fixed number of ions, volume, and temperature. The AIMD simulations were configured to be conducted at temperatures spanning the range from 300 to 600 K, with a time step of 1 fs and a duration of 5 ps. The temperature was maintained using a Nose thermostat model. The transition state search utilizes the fully optimized initial and final structures with RMS convergence set to 0.01 eV Å−1. The activation barriers are calculated from the energy difference between the transition state and the initial state. The calculation of the diffusion potential incorporates corrections for entropy and zero−point energy. The binding energy after Al3+ intercalation into 3R-MoS2 was calculated using the following equation [24]:
E b i n d i n g = E A l x M o S 2 E M o S 2 x E A l x
where E A l x M o S 2 , E M o S 2 , E A l are the total energy of MoS2 with x Al inserted, MoS2 and single Al3+.

4. Conclusions

This paper presents an investigation into the electrochemical properties of 3R-MoS2 and a discussion of its potential use as an AIB. The most stable configuration is one in which a single Al3+ is inserted into a PA octahedron comprising six S atoms. Moreover, the binding energy calculation revealed that the initial Al3+ insertion is random in the ABC stacking structure for 3R-MoS2. The AIMD simulation results indicate that 3R-MoS2 can retain a stable structure within a temperature range of 300 to 600 K. As the number of Al3+ inserted into 3R-MoS2 increases, it is observed that the ions are more likely to be inserted into interlayer sites simultaneously rather than occupying only one interlayer site. The insertion of Al3+ resulted in an increase in the conductivity of 3R-MoS2, accompanied by a transfer of electrons from Al atoms to S atoms, leading to the formation of Al–S ionic bonds. According to the calculation, the theoretical maximum specific capacity of Al3+ intercalated 3R-MoS2 reached 502.30 mAg h−1, and the average voltage of the intercalation was in the range of 0.75−0.96 V. Considering these findings, it can be concluded that 3R-MoS2 has great potential as a cathode material for AIBs.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules29225433/s1, Figure S1: Molecular dynamics simulation analysis at different temperatures as a function of time step and the obtained structures, (a) 300 K, (b) 400 K, (c) 500 K and (d) 600 K; Figure S2: Schematic representations (side view) of the optimized structures of the three different intercalated stages with different Al3+ concentration: 12 (a) 9 (b) 6 (c) 3 (d) Al3+ inserted in 3R-MoS2 for stage-1, 8 (e) 6 (f) 4 (g) 2 (h) Al3+ inserted in 3R-MoS2 for stage-2, 4(i) 3(j) 2(k) 1(l) Al3+ inserted in 3R-MoS2 for stage-3; Figure S3: The 2 × 2 × 2 super-cell of 3R-MoS2 after geometry optimization; Figure S4: Systematic illustration of staging mechanism for (a) 6 Al3+ in stage-2 and stage-1, (b) 4 Al3+ in stage-3 and stage-2, (c) 3 Al3+ in stage-3 and stage-1, (d) 2 Al3+ in stage-3 and stage-2. RE (in eV) is the relative energetics for same concentrations; Figure S5: Binding energy per molecule for all stages of Al3+ as a function weight percentage of Al3+ in AlxMoS2; Figure S6: Total DOSs and partial DOSs of 3R-MoS2 (a) and 12 Al3+ intercalated 3R-MoS2 (b). The Fermi level is set at zero marked in dash line; Figure S7: Hirshfeld charge for stage-1; Figure S8: Convergence tests for the calculation of 3R-MoS2; Table S1: Lattice constants a, c and volume of 3R-MoS2 after geometry optimization compared with other experimental and theoretical data; Table S2: Lattice constants a, c, volume and binding energy of 3R-MoS2 with Al atom inserted in position A (top, middle and bottom position) and B (top, middle and bottom position).

Author Contributions

Conceptualization, B.W.; Methodology, Q.Z. and T.D.; Validation, B.W.; Writing—original draft preparation, B.W.; Writing—review and editing, T.D.; Visualization, R.T., X.L. and C.Z.; Supervision, B.W.; Funding acquisition, B.W. and R.T. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Nature Science Foundation of China (No. U1904198), the Science and Technology Department of Henan Province (No. 222102240071), College Student Innovation Key Project of Henan Province (No. 202311071009), University Key Scientific Research Project of Henan Province (No. 24A480008, No. 24B480014), College Young Teachers Training Program of Henan Province (No. 2023GGJS160), the Fundamental Research Funds for the Central Universities (No. 561224002).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article and Supplementary Materials.

Acknowledgments

The authors gratefully acknowledge HZWTECH for providing computation facilities and Henan Normal University for providing the computational software.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Xu, J.; Cai, X.; Cai, S.; Shao, Y.; Hu, C.; Lu, S.; Ding, S. High-Energy Lithium-Ion Batteries: Recent Progress and a Promising Future in Applications. Energy Environ. Mater. 2023, 6, e12450. [Google Scholar] [CrossRef]
  2. Chang, X.; Zhao, Y.-M.; Yuan, B.; Fan, M.; Meng, Q.; Guo, Y.-G.; Wan, L.-J. Solid-state lithium-ion batteries for grid energy storage: Opportunities and challenges. Sci. China-Chem. 2023, 67, 43–66. [Google Scholar] [CrossRef]
  3. Xie, J.; Lu, Y.-C. A retrospective on lithium-ion batteries. Nat. Commun. 2020, 11, 2499. [Google Scholar] [CrossRef] [PubMed]
  4. Liu, Z.; He, D.; Wang, B.; Wu, T.; Zhao, S.; Li, X.; He, S.; Liang, Y.; Zhou, Y.; Sun, S.; et al. A Low-Voltage Layered Na2TiGeO5 Anode for Lithium-Ion Battery. Small 2022, 18, 2107608. [Google Scholar] [CrossRef] [PubMed]
  5. Langdon, J.; Manthiram, A. A perspective on single-crystal layered oxide cathodes for lithium-ion batteries. Energy Storage Mater. 2021, 37, 143–160. [Google Scholar] [CrossRef]
  6. Wang, C.; Wang, X.; Zhang, R.; Lei, T.; Kisslinger, K.; Xin, H.L. Resolving complex intralayer transition motifs in high-Ni-content layered cathode materials for lithium-ion batteries. Nat. Mater. 2023, 22, 235–241. [Google Scholar] [CrossRef]
  7. Cui, Z.; Xie, Q.; Manthiram, A. A Cobalt- and Manganese-Free High-Nickel Layered Oxide Cathode for Long-Life, Safer Lithium-Ion Batteries. Adv. Energy Mater. 2021, 11, 2102421. [Google Scholar] [CrossRef]
  8. Hudak, N.S. Chloroaluminate-Doped Conducting Polymers as Positive Electrodes in Rechargeable Aluminum Batteries. J. Phys. Chem. C 2014, 118, 5203–5215. [Google Scholar] [CrossRef]
  9. Xu, X.; Hui, K.S.; Hui, K.N.; Shen, J.; Zhou, G.; Liu, J.; Sun, Y. Engineering strategies for low-cost and high-power density aluminum-ion batteries. Chem. Eng. J. 2021, 418, 129385. [Google Scholar] [CrossRef]
  10. Jia, B.; Thang, A.Q.; Yan, C.; Liu, C.; Lv, C.; Zhu, Q.; Xu, J.; Chen, J.; Pan, H.; Yan, Q. Rechargeable Aqueous Aluminum-Ion Battery: Progress and Outlook. Small 2022, 18, 2107773. [Google Scholar] [CrossRef]
  11. Lin, M.-C.; Gong, M.; Lu, B.; Wu, Y.; Wang, D.-Y.; Guan, M.; Angell, M.; Chen, C.; Yang, J.; Hwang, B.-J.; et al. An ultrafast rechargeable aluminium-ion battery. Nature 2015, 520, 324–328. [Google Scholar] [CrossRef] [PubMed]
  12. Zhang, Y.; Liu, S.; Ji, Y.; Ma, J.; Yu, H. Emerging Nonaqueous Aluminum-Ion Batteries: Challenges, Status, and Perspectives. Adv. Mater. 2018, 30, 1706310. [Google Scholar] [CrossRef] [PubMed]
  13. Yang, H.; Li, H.; Li, J.; Sun, Z.; He, K.; Cheng, H.; Li, F. The Rechargeable Aluminum Battery: Opportunities and Challenges. Angew. Chem. Int. Ed. 2019, 58, 11978–11996. [Google Scholar] [CrossRef] [PubMed]
  14. Jayaprakash, N.; Das, S.K.; Archer, L.A. The rechargeable aluminum-ion battery. Chem. Commun. 2011, 47, 12610–12612. [Google Scholar] [CrossRef] [PubMed]
  15. Zhang, Z.; Luo, W.; Liu, Y.; Li, J.; Lu, S.; Chao, Z.; Fan, J. Cubic-like SnO2/ZnS Hollow Heterojunction Encapsulated in Carbon Nanoshell as Cathode for Advanced Aluminum Batteries. Energy Fuels 2024, 38, 1515–1524. [Google Scholar] [CrossRef]
  16. Ou, J.; Luo, W.; Li, J.; Yu, Y.; Zhang, Z.; Chao, Z.; Yi, W.; Fan, J. Sandwich-like CuS@GO as a Cathode for Advanced Aluminum Batteries. Energy Fuels 2024, 38, 6438–6445. [Google Scholar] [CrossRef]
  17. Li, Z.; Niu, B.; Liu, J.; Li, J.; Kang, F. Rechargeable Aluminum-Ion Battery Based on MoS2 Microsphere Cathode. ACS Appl. Mater. Interfaces 2018, 10, 9451–9459. [Google Scholar] [CrossRef]
  18. Tu, J.; Xiao, X.; Wang, M.; Jiao, S. Hierarchical Flower-Like MoS2 Microspheres and Their Efficient Al Storage Properties. J. Phys. Chem. C 2019, 123, 26794–26802. [Google Scholar] [CrossRef]
  19. Tan, D.; Willatzen, M.; Wang, Z.L. Prediction of strong piezoelectricity in 3R-MoS2 multilayer structures. Nano Energy 2019, 56, 512. [Google Scholar] [CrossRef]
  20. Suzuki, R.; Sakano, M.; Zhang, Y.J.; Akashi, R.; Morikawa, D.; Harasawa, A.; Yaji, K.; Kuroda, K.; Miyamoto, K.; Okuda, T.; et al. Valley-dependent spin polarization in bulk MoS2 with broken inversion symmetry. Nat. Nanotechnol. 2014, 9, 611. [Google Scholar] [CrossRef]
  21. Wang, B.; Deng, T.; Liu, J.; Sun, B.; Su, Y.; Ti, R.; Shangguan, L.; Zhang, C.; Tang, Y.; Cheng, N.; et al. Scaly MoS2/rGO Composite as an Anode Material for High-Performance Potassium-Ion Battery. Molecules 2024, 29, 2977. [Google Scholar] [CrossRef] [PubMed]
  22. Wilson, J.A.; Yoffe, A. The transition metal dichalcogenides discussion and interpretation of the observed optical, electrical and structural properties. Adv. Phys. 1969, 18, 193. [Google Scholar] [CrossRef]
  23. Kang, R.; Du, Y.; Zhang, D.; Sun, C.; Zhou, W.; Wang, H.; Chen, G.; Zhang, J. Modification of 2D materials using MoS2 as a model for investigating the Al-storage properties of diverse crystal facets. J. Mater. Chem. A 2023, 11, 15509–15517. [Google Scholar] [CrossRef]
  24. Bhauriyal, P.; Mahata, A.; Pathak, B. The staging mechanism of AlCl4 intercalation in a graphite electrode for an aluminium-ion battery. Phys. Chem. Chem. Phys. 2017, 19, 7980–7989. [Google Scholar] [CrossRef] [PubMed]
  25. Park, J.; Afrinish Fatima, S. A DFT study of TiC3 as anode material for Li-ion batteries. Appl. Surf. Sci. 2023, 638, 158024. [Google Scholar] [CrossRef]
  26. Zhao, Z.; Yu, T.; Zhang, S.; Xu, H.; Yang, G.; Liu, Y. Metallic P3C monolayer as anode for sodium-ion batteries. J. Mater. Chem. A 2018, 7, 405–411. [Google Scholar] [CrossRef]
  27. Raju, V.; Kumar, V.N.Y.; Jetti, V.R.; Basak, P. MoS2/Polythiophene Composite Cathode as a Potential Host for Rechargeable Aluminum Batteries: Deciphering the Impact of Processing on the Performance. ACS Appl. Energy Mater. 2021, 4, 9227–9239. [Google Scholar] [CrossRef]
  28. Tan, B.; Lu, T.; Luo, W.; Chao, Z.; Dong, R.; Fan, J. A Novel MoS2-MXene Composite Cathode for Aluminum-Ion Batteries. Energy Fuels 2021, 35, 12666–12670. [Google Scholar] [CrossRef]
  29. Wu, Y.; Ren, C.; Wei, Q. Novel ternary compound transition metal dichalcogenide TiNbS4 as promising anodes materials for Li-ion batteries: A DFT study. Appl. Surf. Sci. 2023, 615, 156322. [Google Scholar] [CrossRef]
  30. Guo, S.; Yang, H.; Liu, M.; Feng, X.; Xu, H.; Bai, Y.; Wu, C. Interlayer-Expanded MoS2/N-Doped Carbon with Three-Dimensional Hierarchical Architecture as a Cathode Material for High-Performance Aluminum-Ion Batteries. ACS Appl. Energy Mater. 2021, 4, 7064–7072. [Google Scholar] [CrossRef]
  31. Tan, B.; Han, S.; Luo, W.; Chao, Z.; Fan, J.; Wang, M. Synthesis of RGO-supported layered MoS2 with enhanced electrochemical performance for aluminum ion batteries. J. Alloys Compd. 2020, 841, 155732. [Google Scholar] [CrossRef]
  32. Zhao, S.; Li, J.; Chen, H.; Zhang, J. Synthesis of Bi2S3/MoS2 Nanorods and Their Enhanced Electrochemical Performance for Aluminum Ion Batteries. J. Electrochem. Energy Convers. Storage 2020, 17, 031010. [Google Scholar] [CrossRef]
  33. Hohenberg, P.; Kohn, W. Inhomogeneous Electron Gas. Phys. Rev. 1964, 136, B864–B871. [Google Scholar] [CrossRef]
  34. Kohn, W.; Sham, L.J. Self-Consistent Equations Including Exchange and Correlation Effects. Phys. Rev. 1965, 140, A1133–A1138. [Google Scholar] [CrossRef]
  35. Milman, V.; Winkler, B.; White, J.A.; Pickard, C.J.; Payne, M.C.; Akhmatskaya, E.V.; Nobes, R.H. Electronic structure, properties, and phase stability of inorganic crystals: A pseudopotential plane-wave study. Int. J. Quantum Chem. 2000, 77, 895–910. [Google Scholar] [CrossRef]
  36. Blöchl, P.E. Projector augmented-wave method. Phys. Rev. B 1994, 50, 17953–17979. [Google Scholar] [CrossRef]
  37. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758–1775. [Google Scholar] [CrossRef]
  38. Jellinek, F.; Brauer, G.; Müller, H. Molybdenum and Niobium Sulphides. Nature 1960, 185, 376–377. [Google Scholar] [CrossRef]
  39. Schönfeld, B.; Huang, J.J.; Moss, S.C. Anisotropic mean-square displacements (MSD) in single-crystals of 2H- and 3R-MoS2. Acta Crystallogr. B 1983, 39, 404–407. [Google Scholar] [CrossRef]
  40. Coutinho, S.S.; Tavares, M.S.; Barboza, C.A.; Frazao, N.F.; Moreira, E.; Azevedo, D.L. 3R and 2H polytypes of MoS2: DFT and DFPT calculations of structural, optoelectronic, vibrational and thermodynamic properties. J. Phys. Chem. Solids 2017, 111, 25–33. [Google Scholar] [CrossRef]
Figure 1. Two preferred inserted positions of Al3+ in 3R-MoS2 (a) at position A and (b) at position B.
Figure 1. Two preferred inserted positions of Al3+ in 3R-MoS2 (a) at position A and (b) at position B.
Molecules 29 05433 g001
Figure 2. Schematic representations (side view) of the optimized structures of the four different intercalated stages with 0, 12, 8, 4 Al3+ inserted in 3R-MoS2: (a) none, (b) stage-1, (c) stage-2, and (d) stage-3.
Figure 2. Schematic representations (side view) of the optimized structures of the four different intercalated stages with 0, 12, 8, 4 Al3+ inserted in 3R-MoS2: (a) none, (b) stage-1, (c) stage-2, and (d) stage-3.
Molecules 29 05433 g002
Figure 3. Total DOSs and partial DOSs of Al3+ intercalated 3R-MoS2 for stage-1. The Fermi level is set at zero, marked by the dashed line.
Figure 3. Total DOSs and partial DOSs of Al3+ intercalated 3R-MoS2 for stage-1. The Fermi level is set at zero, marked by the dashed line.
Molecules 29 05433 g003
Figure 4. (a) Schematic representation of the diffusion barriers for the three pathways, (b) comparison of diffusion barriers for different paths.
Figure 4. (a) Schematic representation of the diffusion barriers for the three pathways, (b) comparison of diffusion barriers for different paths.
Molecules 29 05433 g004
Figure 5. Voltage profile diagram of Al3+ intercalated 3R-MoS2 system against Al/A3+.
Figure 5. Voltage profile diagram of Al3+ intercalated 3R-MoS2 system against Al/A3+.
Molecules 29 05433 g005
Table 1. Interlayer distance (Ind1, Ind2, Ind3), average open-circuit voltage (OCV), and binding energy (BE) per Al3+ for all stages with different content considered.
Table 1. Interlayer distance (Ind1, Ind2, Ind3), average open-circuit voltage (OCV), and binding energy (BE) per Al3+ for all stages with different content considered.
StagesNo. of AlOCV (V)BE (eV)Ind1 (Å)Ind2 (Å)Ind3 (Å)
Stage-130.96−2.873.163.193.16
60.94−2.853.133.193.31
90.94−2.832.422.513.26
120.77−2.713.232.863.15
Stage-220.94−2.833.152.783.22
40.86−2.693.332.793.23
60.86−2.653.192.783.05
80.81−2.603.172.803.22
Stage-310.92−2.762.783.212.78
20.82−2.612.793.152.75
30.79−2.532.773.212.78
40.75−2.462.813.182.77
Table 2. The calculated theoretical specific capacity compared with other data.
Table 2. The calculated theoretical specific capacity compared with other data.
CathodeTheoretical CapacityInitial Capacity
(mA h g−1)/
Current Density (mA g−1)
Cyclic Capacity (mA h g−1)/
Cycle Number
Discharge Plateau (V)Ref.
stage-1502.300.77–0.96
stage-2334.870.81–0.94
stage-3167.430.75–0.92
2H-MoS2253.6/20 66.7/1000.5–0.9[17]
2H-MoS2/N-doped carbon232/5000.4[30]
2H-MoS2-Maxene224/-166/600.3, 0.9[28]
2H-MoS2/RGO278.1/1000161.1/100[31]
2H-MoS2249.6/100080.9/100[31]
Spongy MoS2214.2/500129.5/10000.5[23]
Bi2S3/MoS2274.3/375132.9/1000.8[32]
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

Wang, B.; Deng, T.; Zhou, Q.; Zhang, C.; Lu, X.; Tao, R. First-Principles Study of 3R-MoS2 for High-Capacity and Stable Aluminum Ion Batteries Cathode Material. Molecules 2024, 29, 5433. https://doi.org/10.3390/molecules29225433

AMA Style

Wang B, Deng T, Zhou Q, Zhang C, Lu X, Tao R. First-Principles Study of 3R-MoS2 for High-Capacity and Stable Aluminum Ion Batteries Cathode Material. Molecules. 2024; 29(22):5433. https://doi.org/10.3390/molecules29225433

Chicago/Turabian Style

Wang, Bin, Tao Deng, Quan Zhou, Chaoyang Zhang, Xingbao Lu, and Renqian Tao. 2024. "First-Principles Study of 3R-MoS2 for High-Capacity and Stable Aluminum Ion Batteries Cathode Material" Molecules 29, no. 22: 5433. https://doi.org/10.3390/molecules29225433

APA Style

Wang, B., Deng, T., Zhou, Q., Zhang, C., Lu, X., & Tao, R. (2024). First-Principles Study of 3R-MoS2 for High-Capacity and Stable Aluminum Ion Batteries Cathode Material. Molecules, 29(22), 5433. https://doi.org/10.3390/molecules29225433

Article Metrics

Back to TopTop