Next Article in Journal
Compressive Failure Characteristics of 3D Four-Directional Braided Composites with Prefabricated Holes
Previous Article in Journal
Regulating TiO2 Deposition Using a Single-Anchored Ligand for High-Efficiency Perovskite Solar Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Theoretical Investigation of a Novel Two-Dimensional Non-MXene Mo3C2 as a Prospective Anode Material for Li- and Na-Ion Batteries

1
School of Physical Science and Technology, Northwestern Polytechnical University, Xi’an 710129, China
2
MSEA International Institute for Materials Genome, Langfang 065500, China
3
Particle Cloud Biotechnology (Hangzhou) Co., Ltd., Hangzhou 310018, China
4
Science and Technology on Thermostructural Composite Materials Laboratory, Northwestern Polytechnical University, Xi’an 710072, China
5
School of Chemistry and Chemical Engineering, Northwestern Polytechnical University, Xi’an 710129, China
*
Authors to whom correspondence should be addressed.
Materials 2024, 17(15), 3819; https://doi.org/10.3390/ma17153819
Submission received: 24 June 2024 / Revised: 17 July 2024 / Accepted: 31 July 2024 / Published: 2 August 2024
(This article belongs to the Special Issue Novel Materials for Electrochemical Energy Storage Systems)

Abstract

:
A new two-dimensional (2D) non-MXene transition metal carbide, Mo3C2, was found using the USPEX code. Comprehensive first-principles calculations show that the Mo3C2 monolayer exhibits thermal, dynamic, and mechanical stability, which can ensure excellent durability in practical applications. The optimized structures of Lix@(3×3)-Mo3C2 (x = 1–36) and Nax@(3×3)-Mo3C2 (x = 1–32) were identified as prospective anode materials. The metallic Mo3C2 sheet exhibits low diffusion barriers of 0.190 eV for Li and 0.118 eV for Na and low average open circuit voltages of 0.31–0.55 V for Li and 0.18–0.48 V for Na. When adsorbing two layers of adatoms, the theoretical energy capacities are 344 and 306 mA h g−1 for Li and Na, respectively, which are comparable to that of commercial graphite. Moreover, the Mo3C2 substrate can maintain structural integrity during the lithiation or sodiation process at high temperature. Considering these features, our proposed Mo3C2 slab is a potential candidate as an anode material for future Li- and Na-ion batteries.

1. Introduction

With the advancement of nanoengineering and microelectronics technology, the exploration of energy-storage devices based on two-dimensional (2D) materials has driven a tremendous amount of research [1]. It is well known that the performance of rechargeable batteries is strongly influenced by the properties of their electrodes. Due to the unique features of 2D materials, such as the large specific surface area, high mechanical strength, high charge carrier capability, and good flexibility [2], rechargeable batteries using 2D materials as electrodes are regarded as a highly promising technology that can effectively integrate clean energy (e.g., solar, wind, and tidal energy) with electricity grids [3]. The ongoing transition from fossil fuels to renewable energy is the key to lowering the CO2 footprint [4]. Currently, the most widely used secondary batteries are lithium-ion batteries (LIBs). Because the shortage of lithium resources is an obstacle to developing LIBs and because sodium is more Earth-abundant and much cheaper than lithium, sodium-ion batteries (SIBs) are a potentially viable battery technology to supplement LIBs [5,6,7].
Among all possible candidates for high-performance electrode materials, MXenes, a large family of 2D transition metal carbides (TMCs) and nitrides (TMNs), are promising candidates for advanced energy-storage devices, such as LIBs, lithium–sulfur batteries, and supercapacitors [1,8,9]. The general formula of MXenes is Mn+1XnTx (n = 1–3), where M donates the transition metal (e.g., V, Mo, Ti, Nb, or Cr), X is nitrogen and/or carbon, and Tx presents the surface termination (e.g., -O, -F, or -OH) [10]. Usually, MXenes can be synthesized by selectively etching the A layer from MAX phases (A represents the group IIIA/IVA element) [10]. Many MXenes have been identified as excellent anodes for secondary batteries, with low diffusion barriers, low operating voltage, and high storage capacity [1,11,12,13]. For example, as an electrode material for LIBs, surface halogenated Ti3C2 MXenes were investigated using first-principles calculations [14]. The results indicated that the Ti3C2T2 MXenes exhibit metallic conductivity with high structural stability and mechanical strength. Compared with Ti3C2F2 and Ti3C2Br2, Ti3C2Cl2 exhibits a large elastic modulus, a low diffusion barrier (0.275 eV), a high open circuit voltage (0.54 eV), and an energy capacity (674.21 mA·h g−1), which is beneficial for high performance as an electrode. Li et al. [15] prepared a Co-doped MXene (Co@MXene) anode for LIBs and the anode exhibits an excellent reversible capacity of 1283.2 mA h g−1 at a current density of 0.1 A g−1 after 120 cycles. Apart from LIBs, a great research effort has also been devoted to the application of MXenes in SIBs [16,17,18]. For example, Fan et al. [19] theoretically studied the properties of the V3C2 MXene as an anode for SIBs and found that the diffusion energy barrier for Na is 0.02 eV, indicating a high rate for charge/discharge processes. The theoretical Na capacity of V3C2 is 606.42 mA h g−1, suggesting a great potential in SIBs. Hence, these research findings indicate that the MXenes possess high potential value in LIBs and SIBs.
It is important to further search for TMC/TMN monolayers that can be used as high-performance electrode materials for rechargeable batteries. Thanks to the considerable progress in the computational algorithms and the computational power of modern computers, the theoretical prediction based on first-principles calculations offers an efficient way to explore energetically stable 2D materials as new electrodes. Xu et al. [20] predicted a new stable 2D TMC containing C2 dimers, VC2, using the swarm-intelligent global-structure search method. As an anode material for LIBs, the VC2 sheet shows a high Li-storage capacity of 1073 mA h g−1 for multilayer adsorption, while stacked VC2 possesses an even larger capacity of 1430 mA h g−1. At the same time, the Li atoms stored in the interlayer of VC2 can migrate easily with a low barrier of 0.09 eV. Using the swarm structural search method, Yu et al. [21] extensively explored Ta-C monolayers with various TaxCy compositions (x = 1 and y = 1–4, or x = 2 and y = 1). They determined the stable, carbon-rich TaC2 as a promising anode for LIBs. The theoretical capacity of the metallic TaC2 reaches 523 mA h g−1 when adsorbing two layers of Li atoms. In addition, the resultant performance of the diffusion barrier and working voltage are better than that of commercial graphite. Wu et al. [22] designed two TMC/TMN monolayers as anodes for LIBs, namely tetr-V2C2 and tetr-V2N2. They reported Li storage capacities of 851 and 824 mA h g−1 for tetr-V2C2 and tetr-V2N2, respectively. These two structures also exhibit low diffusion barriers for Li atoms (smaller than 0.1 eV). Generally, it is very important for the further improvement of rechargeable batteries to predict new types of 2D TMCs/TMNs as potential electrodes.
To our knowledge, so far, only the MO3C2 MXene has been reported [23], with a Mo/C ratio of 3:2. It is, therefore, meaningful to explore new stable structures possessing the same elemental ratio as the Mo3C2 MXene. In this work, a new stable 2D molybdenum carbide, Mo3C2, was predicted using the crystal structure search technique associated with the first-principles calculations. Although the 2D structure has the same chemical formula as the Mo3C2 MXene, its configuration is completely different from the previously reported MXene. The performance of the predicted Mo3C2 as an anode for LIBs/SIBs was comprehensively scrutinized through theoretical calculations.

2. Computational Details

Based on the evolutionary algorithm, the structural prediction of the 2D Mo3C2 was performed with the USPEX code, which allows researchers to efficiently find possible stable structures according to a given chemical composition [24,25]. Specifically, in order to ensure structural diversity in the search process using the USPEX, a large Mo/C elemental ratio of 6:4 was chosen to search low-energy structures, and 60 structures were randomly created in the first generation. The population size of every ensuing generation was set to 60, in which the fractions of structures produced by the heredity, softmutation, transmutation, and random structure generator were 60, 20, 10, and 10%, respectively. The structure relaxation and energy calculation were based on the density functional theory (DFT), as implemented in the VASP 5.4 package [26,27]. The generalized gradient approximation (GGA) [28] in the form of the Perdew–Burke–Ernzerhof (PBE) [29] exchange correction functional was employed to describe the exchange–correlation interactions. For the good structures found with the USPEX, the refined structures and properties were calculated with the VASP package. An energy cutoff of 500 eV was adopted to expand wave functions into plane waves, and the computations were converged within 10−5 eV and 0.02 eV/Å for energy and force criteria, respectively. A vacuum distance of 15 Å (z direction) was used to prevent interaction between adjacent monolayers. The van der Waals interaction was taken into account using the DFT-D3 method [30]. A 3 × 3 supercell was selected to calculate the adsorption and diffusion of Li/Na atoms on the Mo3C2 monolayer. The Brillouin zone was sampled with 10 × 10 × 1 and 5 × 5 × 1 k-point meshes for the unit cell and supercell, respectively. Using the Bader charge method [31], the partial charges of different atoms were calculated, and the amount of charge transfer was estimated quantitatively. The climbing image nudged elastic band (CI-NEB) [32] was employed to investigate the potential migration pathways and energy barriers of adatoms on the substrate.
To examine the dynamical stability of the Mo3C2 monolayer, the phonon frequency was analyzed using the Phonopy code [33]. In addition, ab initio molecular dynamics (AIMD) simulations with the Andersen thermostat [34] and the NVT ensemble were carried out to verify the thermal stability of the studied structures. The size of the supercell of the Mo3C2 was 4 × 5 for the phonon–dispersion relationship calculation and AIMD simulations.

3. Results and Discussion

3.1. Structure and Stability

After the search with the USPEX, a square p4m structure, Mo3C2, was found to be the most energetically stable configuration. Figure 1a,c present the geometric configuration of Mo3C2 associated with the three adsorption sites considered for the Li and Na atoms. The optimized Mo3C2 has the lattice parameters of a = b = 2.954 Å, and the thickness of the monolayer is 4.523 Å. The in-plane Mo-C bond length is 2.117 Å, smaller than the interatomic distance of 2.261 Å in the z direction. It should be noted that the surface of the monolayer is not absolutely planar, with a buckled height of 0.344 Å for each surface. Different from the configuration of the Mo3C2 MXene [23], where every sub-layer contains only one type of element, the surface sub-layers in our predicted Mo3C2 contain both molybdenum and carbon species. It should be noted that, although the Mo atoms in the middle layer of the Mo3C2 slab are only two-fold coordinated with their neighboring C atoms, the large gaps surrounding the Mo atoms can accommodate the electrons of the Mo atoms, resulting in a suitable environment for them. All the C atoms are five-fold coordinated with their neighboring Mo atoms. Similar five-fold coordination can also be found in other stable 2D transition metal carbides, such as TaC [21] and tetr-V2C2 [22]. Notably, all the C atoms are exposed on the surfaces, which can facilitate the adsorption of the Li/Na atoms [35].
Cohesive energy Ecoh refers to the energy required to separate solids into free atoms and can be used to evaluate the bond strength of the structure [36,37]. Ecoh can be calculated as [22]
E coh = [ E ( A m B n ) m E ( A ) n E ( B ) ] / ( m + n )
where E(AmBn), E(A), and E(B) represent the energies per formula unit of AmBn and isolated atoms A and B, respectively. In general, a smaller Ecoh value reflects better stability for the corresponding structure. Here, the calculated Ecoh for the predicted Mo3C2 is −8.45 eV/atom, higher than that of the Mo3C2 MXene (−8.28 eV/atom), suggesting its high energetic stability. For the dynamical stabilities, the phonon spectra of the Mo3C2 have no obvious imaginary mode in the Brillouin zone (Figure 1b), indicating that our predicted Mo3C2 is dynamically stable. In addition, AIMD simulations were used to examine the thermal stability of the predicted Mo3C2 at 600 K with a time step of 3 fs. As illustrated in Figure 1d, the energies and temperatures of the system fluctuated around the equilibrium positions. At the end of the 10 ps simulations, the framework of the Mo3C2 maintained its initial configuration well (Figure S1), demonstrating its high thermal stability at high temperatures.
A promising electrode material should have sufficient in-plane rigidity to resist deformation. The four elastic constants (C11, C22, C12, and C66) [38] and orientation-dependent Young’s moduli (Y) of the predicted Mo3C2 are listed in Table 1. The in-plane 2D Young’s moduli in the Cartesian [01] and [10] directions are [38]
Y [ 01 ] 2 D = C 11 C 22 C 12 2 C 11   and   Y [ 10 ] 2 D = C 11 C 22 C 12 2 C 22
The elastic constants and Y of the 2D Mo3C2 MXene, graphene [38], BN [38], and SiC [38] are included for comparison. All the elastic constants of the predicted Mo3C2 meet the Born–Huang elastic stability criteria, C11C22C122 > 0 and C66 > 0 [39], indicating mechanical stability. Compared with other monolayers, the Y values of our predicted Mo3C2 are much larger than those of the Mo3C2 MXene, BN, and SiC, and slightly higher than those of graphene. All these results suggest that the predicted Mo3C2 is robust enough to form a free-standing membrane.

3.2. Lithiation and Sodiation Processes: Adsorption and Electronic Property

In this work, a 3 × 3 supercell was adopted to study the lithiation and sodiation of the predicted 2D Mo3C2. According to the symmetry of the Mo3C2 monolayer, three representative adsorption sites for M (Li or Na), namely C1 (on-top C), H (4-fold hollow), and Mo1 (on-top Mo), were considered to explore the Li/Na-adsorption properties (Figure 1a). The adsorption energy Ead of M was estimated via the following expression [19]:
E ad = ( E M x @ ( 3 × 3 ) - Mo 3 C 2 E ( 3 × 3 ) - Mo 3 C 2 x E M ) / x
where E M x @ ( 3 × 3 ) - Mo 3 C 2 (M = Li or Na), E ( 3 × 3 ) - Mo 3 C 2 , and E M are the energies of the Li/Na-adsorbed (3×3)-Mo3C2 sheet, the pristine (3×3)-Mo3C2 substrate, and a Li/Na atom in the bulk structure, respectively. In addition, the variable x is the number of adsorbed alkali atoms. For a single adatom, the adsorption energies for all the considered sites are listed in Table 2, and the optimized configurations with the lowest Ead are shown in Figure 2a and Figure 3a. For both Li- and Na-adsorbed systems, the H site is the most favorable adsorption site, as it has the smallest Ead of −0.646 and −0.617 eV for Li and Na, respectively. On the other hand, the Mo1 site has the maximum Ead for all the systems, indicating the weakest adsorption strength. Apart from adsorption energy, the adatom height (h) can also reflect the adsorption strength for different sites. The values of h for single-adatom adsorption on different sites are listed in Table 2. For C1 and Mo sites, h is the vertical distance between the adatom and the surface C or Mo atom directly below the adatom. In contrast, for the H site, h is the distance between the adatom and the center of two surface C atoms around the H site. The H site has the smallest h of 1.66 Å for Li and 2.14 Å for Na, suggesting the strongest adsorption strength. The height at Mo1 is significantly larger than that at C1 and H sites, with h values of 2.47 and 2.76 Å for Li and Na, respectively. These results indicate that the adsorption strength for a single adatom is highest at site H, followed by the C1 site and then the Mo1 site. This strength order is consistent with the corresponding order obtained from the adsorption energy.
As the charges of surface atoms can influence adsorption site preference, Bader charges for the pristine Mo3C2 and single-adatom adsorbed systems were calculated and are listed in Table 3. In the pristine Mo3C2, the C atoms (C1 sites) are negatively charged, leading to an affinity with the electropositive Li/Na atoms. Conversely, the Mo atoms (Mo1 sites) are positively charged, causing weak adsorption for adatoms. In addition, the electron density depletion of 4-fold hollow sites (H sites) allows the localization of alkali metal atoms. To visualize the charge distribution in single-adatom adsorption systems, the differential charge density, ∆ρ, was calculated, and the results are shown in Figure 2. The electrons tend to transfer from the adatoms to the substrate, introducing a built-in electric field to the adsorbed systems. The number of electrons transferred from Li to the substrate is 0.1 e larger than from Na to the substrate (Table 3).
As the electronic structure can influence the battery performance of the anode material, we analyzed the density of states (DOSs) of the Mo3C2 and single-adatom adsorbed Mo3C2, with the adatom located on the stable H site. As shown in Figure S2, a high DOS peak near the Fermi level can be observed for the predicted Mo3C2 monolayer, implying its excellent electronic conductivity. The partial DOSs curves indicate that the DOS at the Fermi level is mainly contributed by the Mo-d and C-p orbitals. There is an obvious overlap near the Fermi level between the Mo-d and C-p orbitals, indicating strong p-d orbital hybridization between the C and Mo species. After adsorbing a metal atom, the adsorbed systems also exhibit metal characteristics (Figure 3). The DOS at the Fermi level is dominated by the Mo-d orbital in the Li@(3×3)-Mo3C2, while in the Na@(3×3)-Mo3C2 it is dominated by the Mo-d and C-p orbitals. It is noteworthy that the systems maintain a metallic nature after adsorption, which can ensure good electronic conduction and is indispensable for an ideal battery electrode.
Next, we investigated the configurations of adsorbing multiple Li/Na atoms. Figure 4 shows the most stable configurations of the Lix@(3×3)-Mo3C2 at x = 1, 6, 12, and 18, and Figure S3 displays the optimal structures at x = 1–18. As the number of adsorbed Li atoms increases, the number of adatoms located on the top of the C atoms (C1 sites) rises gradually, and the C1 sites become the majority at x > 9. In addition, the Li atoms cannot be stabilized on top of the Mo atoms. For the saturated concentration of the first layer on the (3×3)-Mo3C2, the structure with the lowest energy contains 6 Li atoms on the H sites and 12 Li atoms on the C1 sites. For the Na-adsorbed systems, Figure 5 shows the most stable configurations of the Nax@(3×3)-Mo3C2 at x = 1, 6, 12, and 18, and Figure S5 presents every optimized structure at x = 1–18. Similarly to the Li-adsorbed systems, the most common adsorption site types for Na are C1 and H, and the Na18@(3×3)-Mo3C2 contains 6 Na atoms on the H sites and 12 Na atoms on the C1 sites. It is noteworthy that the adsorption energy of the optimal configurations for the saturated concentration is 0.023 eV for Li and 0.010 eV for Na, lower than the configurations where all the adatoms are adsorbed on the C1 sites. As shown in Figure 4b and Figure 5b, for x = 6, multiple Li atoms can be found on both sides of the substrate, while the Na atoms are mostly adsorbed on one surface. This is related to the fact that the substrate and previous adatoms have a joint impact on the adsorption positions of subsequent adatoms. On the one hand, the substrate and previous adatoms have an adsorption effect for the following adatoms. On the other hand, the accumulation of charge transferred from the alkali atoms on the surface reduces the adsorption effect. Different from the first adatom, the second Li/Na atom tends to be adsorbed on another surface due to weak attraction from the first adatom. Due to the Bader charges of different atoms, as shown in Table 3, Li transfers more charge to the substrate than Na atoms. Compared with the Na counterpart, in the Li-adsorbed system, a greater charge accumulation on the surface of the substrate will result in a weaker attraction for subsequent adatoms. This results in the adsorption of multiple Li atoms on both sides of the Mo3C2 slab (Figure S3). For the Na-adsorbed system, when the number of adatoms ranges from 4 to 7, the adsorption effect from previous adatoms slightly decreases due to a small amount of charge accumulation, resulting in most of the Na atoms being adsorbed on the same surface (Figure S5). When the number of Na adatoms is greater than 7, the charge accumulation is big enough to result in adsorption on both surfaces. For one-layer adsorption, the lattice constants increase from 8.834 Å for the (3×3)-Mo3C2 to 8.886 Å (about 0.59% tensile strain) for the Li18@(3×3)-Mo3C2 and to 8.924 Å (about 1.02% tensile strain) for the Na18@(3×3)-Mo3C2. The small expansions in the Li/Na adsorption processes suggest excellent substrate stability.
The adsorption of the second layer of Li/Na atoms on the Mo3C2 substrate with the first layer of adatoms was also studied, and the corresponding configurations are displayed in Figures S4 and S6 for Li and Na, respectively. The results show that the Li atoms tend to be located on the hollow sites or near the top of the Mo atoms. A maximum of 18 Li atoms are deposited on the Li18@(3×3)-Mo3C2, resulting in two Li monolayers per substrate surface. With regard to the Na atoms on the Na18@(3×3)-Mo3C2, the majority tend to migrate to the positions near Mo1 sites. The maximum number of Na atoms in the second layer is 14, less than that of Li atoms in the Li-adsorbed system. This difference is due to the larger radius of sodium compared to lithium. For double-layer adsorption, the lattice constants are 8.864 Å for the Li36@(3×3)-Mo3C2 and 8.928 Å for the Na32@(3×3)-Mo3C2. It is worth noting that the lattice constant for the double-layer adsorption of lithium atoms is smaller than for the single-layer adsorption of lithium atoms. This may be due to the strong interaction between the Li atoms. Generally, the adsorption of Li/Na atoms has little effect on the lattice expansion of the substrate, which is beneficial for improving the life span of the electrode.

3.3. Li/Na Surface Diffusion

It is believed that the diffusion barriers of adatoms on the surface of the substrate influence the charge/discharge rate of the electrode. In this work, as shown in Figure 6a,c, two diffusion pathways between two adjacent H sites were considered and labeled path C1 (through the H-C1-H line) and path Mo1 (through the H-Mo1-H line). The CI-NEB [31] method was applied, and nine images were adopted to calculate the migration energy barriers. For the Li-adsorbed system, the C1 pathway has the lowest energy barrier of 0.190 eV (Figure 6b), which is much smaller than for path Mo1 (0.508 eV). In the case of Na diffusion, the adatom tends to migrate through path C1 with a barrier of 0.118 eV, which is similar to Li diffusion. The activation barrier for path Mo1 (0.169 eV) is approximately three times smaller than for Li. Compared with the Li atom, the larger distance between the Na atom and the surface of the Mo3C2 sheet can effectively reduce the diffusion barriers for Na. Therefore, the diffusion barriers for Li are higher than for Na overall. The lowest barriers for Li/Na on the Mo3C2 are remarkably smaller than those on typical 2D materials, such as graphene (0.31 eV for Li [40] and 0.18 eV for Na [41]) and silicene (0.28 eV for Li [42] and 0.25 eV for Na [43]). These low energy barriers are highly beneficial to applying the predicted Mo3C2 in LIBs and SIBs.

3.4. Theoretical Storage Capacity, Open-Circuit Voltage, and Thermal Stability

The specific energy capacity of anode material is proportional to the number of adsorbed adatoms. The storage capacity of the predicted 2D Mo3C2 host with two layers of adatoms was investigated in this work. The maximum theoretical capacity (CM) can be obtained using the following formula [19]:
C M = z w max F / M Mo 3 C 2
where z is the number of valence electrons of the adatom, wmax is the maximum adatom concentration, F is Faraday’s constant (26,801 mA h mol−1), and M Mo 3 C 2 is the molar mass of the Mo3C2. After adsorbing two layers of adatoms, the Mo3C2 can adsorb up to 36 Li atoms and 32 Na atoms, corresponding to the concentrations of 4 and 3.56 for Li and Na, respectively. The maximum theoretical storage capacities of the Mo3C2 were calculated as 344 mA h g−1 for Li and 306 mA h g−1 for Na, which are comparable to that of commercial graphite (372 mA h g−1) [44]. The results suggest that the 2D Mo3C2 can serve as an anode material for LIBs and SIBs.
The charge/discharge processes of the Mo3C2 sheet can be described by the following half-cell reaction [19]:
Mo3C2 + wM+ + we ↔ Mo3C2Mw
The open-circuit voltage (OCV) is estimated via the following equation [45]:
OCV ( E Mo 3 C 2 + w E M E Mo 3 C 2 M w ) / w z e
The calculated OCV values, as a function of the concentration of adatoms on the Mo3C2 supercell, are presented in Figure 7a. The highest voltages for Li and Na are 0.65 and 0.62 V, respectively, corresponding to the single adatom on the Mo3C2 supercell. For both Li- and Na-adsorption systems, the OCV curves decrease initially and tend to be relatively moderate in the following stage. Subsequently, the voltage curves once again display a downward trend from concentrations w = 2 for Li and w = 1.33 for Na. In addition, Figure 7b displays the average OCV values of Mw@2D-Mo3C2 for four regions of the entire concentration range. The average OCV values for all the systems generally decline as the concentration of adatoms increases, with ranges of 0.31 to 0.55 V and of 0.18 to 0.48 V for Liw@2D-Mo3C2 and Naw@2D-Mo3C2, respectively. Generally, it is accepted that the OCV values for anode materials are between 0.1 and 1.0 V [22]. Thus, the OCV results suggest that the 2D Mo3C2 is a potential candidate as an anode material for LIBs and SIBs.
AIMD simulations for the Li18@(3×3)-Mo3C2 and Na18@(3×3)-Mo3C2 at temperatures of 300 and 600 K were carried out with a time step of 3 fs to study the reliability of the anode material under different temperatures. The energy fluctuations and final configurations are presented in Figure 8. The energies of the systems do not display big fluctuations around the equilibrium position. For AIMD simulations at 300 K, the structures exhibit no significant deformation. In the case of 600 K, although there is a significant change in the position of adsorbed metal atoms, the Mo3C2 substrate exhibits no obvious deformation. Overall, the predicted Mo3C2 slab possesses good thermal stability as an anode for LIBs and SIBs.

4. Summary

In summary, in this study, we found a stable 2D non-MXene structure, Mo3C2, using the USPEX technique. First-principles calculations demonstrated that the Mo3C2 monolayer not only is thermally, dynamically, and mechanically stable but also exhibits metallic behavior. As an anode material for LIBs and SIBs, the Mo3C2 possesses the good properties of low diffusion barriers (0.190 eV for Li and 0.118 for Na) and low electrode potentials. The storage capacities are 344 mA h g−1 for Li and 306 mA h g−1 for Na when adsorbing two layers of adatoms, which are comparable to those of conventional graphite. Moreover, the Mo3C3 substrate can maintain structural integrity during the AIMD simulations for the Li18@(3×3)-Mo3C2 and Na18@(3×3)-Mo3C2 at 300 and 600 K. All these results support the idea that the 2D Mo3C2 monolayer can serve as an anode material for LIBs and SIBs.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ma17153819/s1, Figure S1. (a) Top and (b) side views of the Mo3C2 monolayer after 10 ps AIMD simulations at 600 K. Figure S2. Total and partial densities of states of the Mo3C2 slab. The fermi level is set to zero and marked with the dashed line. Figure S3. Top and side views of the supercells for the most stable configuration of the single-layer 1–18 Li atom(s) adsorbed on the Mo3C2 surface(s). Figure S4. Top and side views of the supercells for the most stable configuration of the double-layer 19–36 Li atoms adsorbed on the Mo3C2 surfaces. Figure S5. Top and side views of the supercells for the most stable configuration of the single-layer 1–18 Na atom(s) adsorbed on the Mo3C2 surface(s). Figure S6. Top and side views of the supercells for the most stable configuration of the double-layer 19–32 Na atoms adsorbed on the Mo3C2 surfaces.

Author Contributions

Conceptualization, B.X.; Methodology, S.Y.; Software, Q.Z. and S.Y.; Formal analysis, K.S.; Resources, Q.Z. and K.S.; Writing—original draft, B.X.; Visualization, B.X.; Supervision, K.S.; Project administration, K.S.; Funding acquisition, K.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China (grant no. 51761135032), the MSEA International Institute for Materials Genome, and the high-performance computing center of Northwestern Polytechnical University.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the article/Supplementary Materials, further inquiries can be directed to the corresponding authors.

Conflicts of Interest

Authors Qingfeng Zeng and Shuyin Yu were employed by the company Particle Cloud Biotechnology (Hangzhou) Co., Ltd. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Wang, C.; Guan, S.; Zhang, H.; Shen, R.; Yuan, H.; Li, B. Perspectives on two-dimensional ultra-thin materials in energy catalysis and storage. APL Mater. 2023, 11, 050902. [Google Scholar] [CrossRef]
  2. Baig, N. Two-dimensional nanomaterials: A critical review of recent progress, properties, applications, and future directions. Compos. Part A-Appl. S. 2023, 165, 107362. [Google Scholar] [CrossRef]
  3. Hu, Z.; Liu, Q.; Chou, S.L.; Dou, S.X. Two-dimensional material-based heterostructures for rechargeable batteries. Cell Rep. Phys. Sci. 2021, 2, 100286. [Google Scholar] [CrossRef]
  4. Tarascon, J.M. Na-ion versus Li-ion batteries: Complementarity rather than competitiveness. Joule 2020, 4, 1616–1620. [Google Scholar] [CrossRef]
  5. Abraham, K.M. How comparable are sodium-ion batteries to lithium-ion counterparts? ACS Energy Lett. 2020, 5, 3544–3547. [Google Scholar] [CrossRef]
  6. Zhao, D.; Jiang, S.; Yu, S.; Ren, J.; Zhang, Z.; Liu, S.; Liu, X.; Wang, Z.; Wu, Y.; Zhang, Y. Lychee seed-derived microporous carbon for high-performance sodium-sulfur batteries. Carbon 2023, 201, 864–870. [Google Scholar] [CrossRef]
  7. Zhang, Y.; Zhang, Z.; Yu, S.; Johnson, H.M.; Zhao, D.C.; Tan, S.C.; Pan, Z.D.; Wang, Z.L.; Wu, Y.T.; Liu, X. Three-dimensional nanostructured Co2VO4-decorated carbon nanotubes for sodium-ion battery anode materials. Rare Met. 2023, 42, 4060–4069. [Google Scholar] [CrossRef]
  8. Fan, K.; Huang, H. Two-dimensional host materials for lithium-sulfur batteries: A review and perspective. Energy Storage Mater. 2022, 50, 696–717. [Google Scholar] [CrossRef]
  9. Forouzandeh, P.; Pillai, S.C. Two-dimensional (2D) electrode materials for supercapacitors. Mater. Today Proc. 2021, 41, 498–505. [Google Scholar] [CrossRef]
  10. Naguib, M.; Mochalin, V.N.; Barsoum, M.W.; Gogotsi, Y. 25th anniversary article: MXenes: A new family of two-dimensional materials. Adv. Mater. 2014, 19, 992–1005. [Google Scholar] [CrossRef]
  11. Zhang, X.; Zhang, Z.; Zhou, Z. MXene-based materials for electrochemical energy storage. J. Energy Chem. 2018, 27, 73–85. [Google Scholar] [CrossRef]
  12. Shetti, N.P.; Mishra, A.; Basu, S.; Aminabhavi, T.M.; Alodhayb, A.; Pandiaraj, S. MXenes as Li-ion battery electrodes: Progress and outlook. Energy Fuels 2023, 37, 12541–12557. [Google Scholar] [CrossRef]
  13. Wu, Y.; Sun, Y.; Zheng, J.; Rong, J.; Li, H.; Niu, L. MXenes: Advanced materials in potassium ion batteries. Chem. Eng. J. 2021, 404, 126565. [Google Scholar] [CrossRef]
  14. Xiao, M.; Li, M.; Song, E.; Song, H.; Li, Z.; Bi, J. Halogenated Ti3C2 MXene as high capacity electrode material for Li-ion batteries. J. Inorg. Mater. 2022, 37, 660–668. [Google Scholar] [CrossRef]
  15. Li, Z.; Xiao, X.; Zhang, D.; Niu, H.; Shu, Q.; Yan, C.; Wang, X.; Guan, P.; Hu, X. Electrostatics and chemistry combination divalent cobalt ions and alkali-treated MXene for high-performance lithium-ion batteries. Energy Fuels 2022, 36, 13266–13277. [Google Scholar] [CrossRef]
  16. Li, J.; Liu, H.; Shi, X.; Li, X.; Li, W.; Guan, E.; Lu, T.; Pan, L. MXene-based anode materials for high performance sodium-ion batteries. J. Colloid Interf. Sci. 2024, 658, 425–440. [Google Scholar] [CrossRef]
  17. Wu, Y.; Nie, P.; Wang, J.; Dou, H.; Zhang, X. Few-layer MXenes delaminated via high-energy mechanical milling for enhanced sodium-ion batteries performance. ACS Appl. Mater. Interfaces 2017, 9, 39610–39617. [Google Scholar] [CrossRef] [PubMed]
  18. Zhu, J.; Wang, M.; Lyu, M.; Jiao, Y.; Du, A.; Luo, B.; Gentle, I.; Wang, L. Two-dimensional titanium carbonitride Mxene for high-performance sodium ion batteries. ACS Appl. Nano Mater. 2018, 1, 6854–6863. [Google Scholar] [CrossRef]
  19. Fan, K.; Ying, Y.; Li, X.; Luo, X.; Huang, H. Theoretical investigation of V3C2 MXene as prospective high-capacity anode material for metal-ion (Li, Na, K, and Ca) batteries. J. Phys. Chem. C 2019, 123, 18207–18214. [Google Scholar] [CrossRef]
  20. Xu, J.; Wang, D.; Lian, R.; Gao, X.; Liu, Y.; Yury, G.; Chen, G.; Wei, Y. Structural prediction and multilayer Li+ storage in two-dimensional VC2 carbide studied by first-principles calculations. J. Mater. Chem. A 2019, 7, 8873–8881. [Google Scholar] [CrossRef]
  21. Yu, T.; Zhang, S.; Li, F.; Zhao, Z.; Liu, L.; Xu, H.; Yang, G. Stable and metallic two-dimensional TaC2 as an anode material for lithium-ion battery. J. Mater. Chem. A 2017, 5, 18698–18706. [Google Scholar] [CrossRef]
  22. Wu, Y.Y.; Bo, T.; Zhang, J.; Lu, Z.; Wang, Z.; Li, Y.; Wang, B.T. Novel two-dimensional tetragonal vanadium carbides and nitrides as promising materials for Li-ion batteries. Phys. Chem. Chem. Phys. 2019, 21, 19513–19520. [Google Scholar] [CrossRef] [PubMed]
  23. Li, N.; Chen, X.; Ong, W.J.; MacFarlane, D.R.; Zhao, X.; Cheetham, A.K.; Sun, C. Understanding of electrochemical mechanisms for CO2 capture and conversion into hydrocarbon fuels in transition-metal carbides (MXenes). ACS Nano 2017, 11, 10825–10833. [Google Scholar] [CrossRef]
  24. Oganov, A.R.; Glass, C.W. Crystal structure prediction using ab initio evolutionary techniques: Principles and applications. J. Chem. Phys. 2006, 124, 244704. [Google Scholar] [CrossRef]
  25. Oganov, A.R.; Ma, Y.; Lyakhov, A.O.; Valle, M.; Gatti, C. Evolutionary crystal structure prediction as a method for the discovery of minerals and materials. Rev. Miner. Geochem. 2010, 71, 271–298. [Google Scholar] [CrossRef]
  26. Kresse, G.; Furthmüller, J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 1996, 6, 15–50. [Google Scholar] [CrossRef]
  27. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169–11186. [Google Scholar] [CrossRef]
  28. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [PubMed]
  29. Perdew, J.P.; Ernzerhof, M.; Burke, K. Rationale for mixing exact exchange with density functional approximations. J. Chem. Phys. 1996, 105, 9982–9985. [Google Scholar] [CrossRef]
  30. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104. [Google Scholar] [CrossRef]
  31. Tang, W.; Sanville, E.; Henkelman, G. A grid-based Bader analysis algorithm without lattice bias. J. Phys. Condens. Matter. 2009, 21, 084204. [Google Scholar] [CrossRef] [PubMed]
  32. Henkelman, G.; Jónsson, H. Improved tangent estimate in the nudged elastic band method for finding minimum energy paths and saddle points. J. Chem. Phys. 2000, 113, 9978–9985. [Google Scholar] [CrossRef]
  33. Togo, A.; Tanaka, I. First principles phonon calculations in materials science. Scr. Mater. 2015, 108, 1–5. [Google Scholar] [CrossRef]
  34. Andersen, H.C. Molecular dynamics simulations at constant pressure and/or temperature. J. Chem. Phys. 1980, 72, 2384–2393. [Google Scholar] [CrossRef]
  35. Zhang, F.; Jing, T.; Cai, S.; Deng, M.; Liang, D.; Qi, X. Two-dimensional ZrC2 as a novel anode material with high capacity for sodium ion battery. Phys. Chem. Chem. Phys. 2021, 23, 12731–12738. [Google Scholar] [CrossRef] [PubMed]
  36. Singh, M.; Phantsi, T.D. Bond theory model to study cohesive energy, thermal expansion coefficient and specific heat of nanosolids. Chin. J. Phys. 2018, 56, 2948–2957. [Google Scholar] [CrossRef]
  37. Mardiyah, R.U.; Arkundato, A.; Misto; Purwandari, E. Energy cohesive calculation for some pure metals using the Lennard-Jones potential in Lammps Molecular Dynamics. J. Phys. Conf. Ser. 2020, 1491, 012020. [Google Scholar] [CrossRef]
  38. Andrew, R.C.; Mapasha, R.E.; Ukpong, A.M.; Chetty, N. Mechanical properties of graphene and boronitrene. Phys. Rev. B 2012, 85, 125428. [Google Scholar] [CrossRef]
  39. Mouhat, F.; Coudert, F.X. Necessary and sufficient elastic stability conditions in various crystal systems. Phys. Rev. B 2014, 90, 224104. [Google Scholar] [CrossRef]
  40. Fan, X.; Zheng, W.T.; Kuo, J.L. Adsorption and diffusion of Li on pristine and defective graphene. ACS Appl. Mater. Interfaces 2012, 4, 2432–2438. [Google Scholar] [CrossRef]
  41. Singh, T.; Choudhuri, J.R.; Rana, M.K. α-graphyne as a promising anode material for Na-ion batteries: A first-principles study. Nanotechnology 2023, 34, 045404. [Google Scholar] [CrossRef] [PubMed]
  42. Setiadi, J.; Arnold, M.D.; Ford, M.J. Li-Ion adsorption and diffusion on two-dimensional silicon with defects: A first principles study. ACS Appl. Mater. Interfaces 2013, 5, 10690–10695. [Google Scholar] [CrossRef] [PubMed]
  43. Guo, X.X.; Guo, P.; Zheng, J.M.; Cao, L.K.; Zhao, P.J. First-principles calculations study of Na adsorbed on silicene. Appl. Surf. Sci. 2015, 341, 69–74. [Google Scholar] [CrossRef]
  44. Zhao, L.; Ding, B.; Qin, X.Y.; Wang, Z.; Lv, W.; He, Y.B.; Yang, Q.H.; Kang, F. Revisiting the roles of natural graphite in ongoing lithium-ion batteries. Adv. Mater. 2022, 34, 2106704. [Google Scholar] [CrossRef] [PubMed]
  45. Zhang, X.; Yu, Z.; Wang, S.S.; Guan, S.; Yang, H.Y.; Yao, Y.; Yang, S.A. Theoretical prediction of MoN2 monolayer as a high capacity electrode material for metal ion batteries. J. Mater. Chem. A 2016, 4, 15224–15231. [Google Scholar] [CrossRef]
Figure 1. (a) Top and (c) side views of the 2D p4m Mo3C2 monolayer, where C1 (top of C), H (hollow), and Mo1 (top of Mo) represent three possible adsorption sites for adatoms, and the dashed and solid lines represent the unit cell and the 3 × 3 supercell, respectively. (b) Phonon dispersion spectra of the Mo3C2. (d) Energy and temperature as a function of time for Mo3C2 during the AIMD simulations.
Figure 1. (a) Top and (c) side views of the 2D p4m Mo3C2 monolayer, where C1 (top of C), H (hollow), and Mo1 (top of Mo) represent three possible adsorption sites for adatoms, and the dashed and solid lines represent the unit cell and the 3 × 3 supercell, respectively. (b) Phonon dispersion spectra of the Mo3C2. (d) Energy and temperature as a function of time for Mo3C2 during the AIMD simulations.
Materials 17 03819 g001
Figure 2. Differential charge density distributions for the H site (hollow 4-fold site): (a,c) Li@(3×3)-Mo3C2; (b,d) Na@(3×3)-Mo3C2 plotted with isovalue = 8 × 10−4 electrons/Bohr3. Yellow denotes charge accumulation, and blue represents depletion.
Figure 2. Differential charge density distributions for the H site (hollow 4-fold site): (a,c) Li@(3×3)-Mo3C2; (b,d) Na@(3×3)-Mo3C2 plotted with isovalue = 8 × 10−4 electrons/Bohr3. Yellow denotes charge accumulation, and blue represents depletion.
Materials 17 03819 g002
Figure 3. Total and partial densities of states of (a) Li@(3×3)-Mo3C2 and (b) Na@(3×3)-Mo3C2 with the adatom at the H site. The Fermi level is set to zero and marked with the dashed line.
Figure 3. Total and partial densities of states of (a) Li@(3×3)-Mo3C2 and (b) Na@(3×3)-Mo3C2 with the adatom at the H site. The Fermi level is set to zero and marked with the dashed line.
Materials 17 03819 g003
Figure 4. Top and side views of the 3×3 supercells for the most stable configuration of Lix@(3×3)-Mo3C2 at x = (a) 1, (b) 6, (c) 12, and (d) 18.
Figure 4. Top and side views of the 3×3 supercells for the most stable configuration of Lix@(3×3)-Mo3C2 at x = (a) 1, (b) 6, (c) 12, and (d) 18.
Materials 17 03819 g004
Figure 5. Top and side views of the 3 × 3 supercells for the most stable configuration of Nax@(3×3)-Mo3C2 at x = (a) 1, (b) 6, (c) 12, and (d) 18.
Figure 5. Top and side views of the 3 × 3 supercells for the most stable configuration of Nax@(3×3)-Mo3C2 at x = (a) 1, (b) 6, (c) 12, and (d) 18.
Materials 17 03819 g005
Figure 6. The migration pathways of (a) Li and (c) Na on the Mo3C2 monolayer. The energy profiles of (b) Li and (d) Na diffusion on the Mo3C2 monolayer.
Figure 6. The migration pathways of (a) Li and (c) Na on the Mo3C2 monolayer. The energy profiles of (b) Li and (d) Na diffusion on the Mo3C2 monolayer.
Materials 17 03819 g006
Figure 7. (a) Open-circuit voltage and (b) average open-circuit voltage as a function of the concentration of adatoms in Mw@2D-Mo3C2.
Figure 7. (a) Open-circuit voltage and (b) average open-circuit voltage as a function of the concentration of adatoms in Mw@2D-Mo3C2.
Materials 17 03819 g007
Figure 8. Energies as a function of time of (a) Li18@(3×3)-Mo3C2 at 300 K, (b) Li18@(3×3)-Mo3C2 at 600 K, (c) Na18@(3×3)-Mo3C2 at 300 K, and (d) Na18@(3×3)-Mo3C2 at 600 K during the AIMD simulations (inset: the structures after 5 ps AIMD simulations).
Figure 8. Energies as a function of time of (a) Li18@(3×3)-Mo3C2 at 300 K, (b) Li18@(3×3)-Mo3C2 at 600 K, (c) Na18@(3×3)-Mo3C2 at 300 K, and (d) Na18@(3×3)-Mo3C2 at 600 K during the AIMD simulations (inset: the structures after 5 ps AIMD simulations).
Materials 17 03819 g008
Table 1. The elastic constants Cij (N/m) and Young’s moduli Y2D (N/m) of our predicted Mo3C2, the Mo3C2 MXene, graphene [38], BN [38], and SiC [38].
Table 1. The elastic constants Cij (N/m) and Young’s moduli Y2D (N/m) of our predicted Mo3C2, the Mo3C2 MXene, graphene [38], BN [38], and SiC [38].
C11C22C12C66 Y [ 01 ] 2 D Y [ 10 ] 2 D
Our predicted Mo3C2432.3432.3168.4195.6366.7366.7
Mo3C2 MXene293.3293.359.7116.8281.1281.1
Graphene352.7352.760.9145.9342.2342.2
BN289.8289.863.7113.1275.8275.8
SiC179.7179.753.962.9163.5163.5
Table 2. Adsorption energies and adatom heights for single Li/Na atom adsorption on the (3×3)-Mo3C2 monolayer.
Table 2. Adsorption energies and adatom heights for single Li/Na atom adsorption on the (3×3)-Mo3C2 monolayer.
Ead (eV)h (Å)
C1HMo1C1HMo1
Li@(3×3)-Mo3C2−0.467−0.646−0.1481.961.662.47
Na@(3×3)-Mo3C2−0.503−0.617−0.4532.352.142.76
Table 3. Calculated Bader atomic charges in the Mo3C2 and (Li or Na)@ (3×3)-Mo3C2 systems.
Table 3. Calculated Bader atomic charges in the Mo3C2 and (Li or Na)@ (3×3)-Mo3C2 systems.
Average Charge
MoCLiNa
Mo3C2+0.7−1.1
Li@(3×3)-Mo3C2 for C1+0.7−1.1+0.9
Li@(3×3)-Mo3C2 for H+0.7−1.2+0.9
Li@(3×3)-Mo3C2 for Mo1+0.7−1.1+0.9
Na@(3×3)-Mo3C2 for C1+0.7−1.1 +0.8
Na@(3×3)-Mo3C2 for H+0.7−1.1 +0.8
Na@(3×3)-Mo3C2 for Mo1+0.7−1.1 +0.8
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

Xue, B.; Zeng, Q.; Yu, S.; Su, K. Theoretical Investigation of a Novel Two-Dimensional Non-MXene Mo3C2 as a Prospective Anode Material for Li- and Na-Ion Batteries. Materials 2024, 17, 3819. https://doi.org/10.3390/ma17153819

AMA Style

Xue B, Zeng Q, Yu S, Su K. Theoretical Investigation of a Novel Two-Dimensional Non-MXene Mo3C2 as a Prospective Anode Material for Li- and Na-Ion Batteries. Materials. 2024; 17(15):3819. https://doi.org/10.3390/ma17153819

Chicago/Turabian Style

Xue, Bo, Qingfeng Zeng, Shuyin Yu, and Kehe Su. 2024. "Theoretical Investigation of a Novel Two-Dimensional Non-MXene Mo3C2 as a Prospective Anode Material for Li- and Na-Ion Batteries" Materials 17, no. 15: 3819. https://doi.org/10.3390/ma17153819

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