Next Article in Journal
Solution-Processed Two-Dimensional Materials for Device Applications
Next Article in Special Issue
Orientation-Dependent Nanomechanical Behavior of Pentaerythritol Tetranitrate as Probed by Multiple Nanoindentation Tip Geometries
Previous Article in Journal
Effect of Ultrasonic Vibration and Average Grain Size on the Deformability of T2 Copper in T-Shaped Micro-Upsetting
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Research on the Mechanical and Photoelectric Properties Regulation of the New-Type Ceramic Material Ta2AlC

School of Electronic and Information Engineering, Anshun University, Anshun 561000, China
*
Author to whom correspondence should be addressed.
Crystals 2025, 15(4), 309; https://doi.org/10.3390/cryst15040309
Submission received: 28 February 2025 / Revised: 22 March 2025 / Accepted: 24 March 2025 / Published: 26 March 2025
(This article belongs to the Special Issue Microstructure and Characterization of Crystalline Materials)

Abstract

:
Ta2AlC is an emerging ceramic material characterized by its high melting point, high hardness, excellent thermal stability, and superior mechanical properties, which allow for broad application prospects in aerospace and defense fields. This paper investigates the physical mechanisms underlying the modulation of the mechanical and photoelectric properties of Ta2AlC through doping using the first-principles pseudopotential plane-wave method. We specifically calculated the geometric structure, mechanical properties, electronic structure, Mulliken population analysis, and optical properties of Ta2AlC doped with V, Ga, or Si. The results indicate that doping induces significant changes in the structural parameters of Ta2AlC. By applying the Born’s criterion as the standard for mechanical stability, we have calculated that the structures of Ta2AlC, both before and after doping, are stable. The mechanical property calculations revealed that V and Si doping weaken the material’s resistance to deformation while enhancing its plasticity. In contrast, Ga doping increases the material’s resistance to lateral deformation and brittleness. Doping also increases the anisotropy of Ta2AlC. Electronic structure calculations confirmed that Ta2AlC is a conductor with excellent electrical conductivity, which is not diminished by doping. The symmetric distribution of spin-up and spin-down electronic state densities indicates that the Ta2AlC system remains non-magnetic after doping. The partial density of states diagrams successfully elucidated the influence of dopant atoms on the band structure and electronic state density. Mulliken population analysis revealed that V and Ga doping enhance the covalent interactions between C-Ta and Al-Ta atoms, whereas Si doping weakens these interactions. Optical property calculations showed that V and Si doping significantly enhance the electromagnetic energy storage capacity and dielectric loss of Ta2AlC, while Ga doping has minimal effect. The reflectivity of doped and undoped Ta2AlC reaches over 90% in the ultraviolet region, indicating its potential as an anti-ultraviolet coating material. In the visible light region, both doped and undoped Ta2AlC exhibit a similar metallic gray appearance, suggesting its potential as a temperature control coating material. The light loss of Ta2AlC is limited to a narrow energy range, indicating that doping does not affect its use as a light storage material. These results demonstrate that different dopants can effectively modulate the mechanical and photoelectric properties of Ta2AlC.

1. Introduction

With the continuous advancement of technology, the requirements for material properties have become increasingly stringent. Traditional ceramic materials, due to their relatively low strength and toughness, face certain limitations regarding their areas of application. To overcome these shortcomings, a new type of ternary layered ceramic material has emerged, with the general formula Mn+1AXn (where n = 1, 2, 3), commonly referred to as MAX phase ceramic materials [1,2]. In these materials, M represents transition metal elements, including titanium (Ti), vanadium (V), tantalum (Ta), and zirconium (Zr), among others; A denotes elements from group IIIA or ⅣA, such as aluminum (Al), silicon (Si), gallium (Ga), or germanium (Ge); and X is typically the element carbon (C) or nitrogen (N). MAX phase ceramic materials integrate the excellent properties of both metals and ceramics. At room temperature, they exhibit not only the superior electrical and thermal conductivity of metals but also relatively low Vickers microhardness and high values of elastic and shear moduli [3,4,5,6]. These materials can be mechanically processed like metals and maintain good plasticity at elevated temperatures. Moreover, they possess the high yield strength, high melting point, high thermal stability, and good oxidation resistance characteristics of ceramic materials [7,8,9,10,11,12,13,14,15,16,17]. Notably, this new type of ceramic demonstrates excellent strength and a very low friction coefficient at high temperatures, and like graphite, it also has good self-lubricating properties. These attributes render MAX phase ceramic materials a new type of structural and functional integrated ceramic material with broad potential for application.
The properties of MAX phase ceramic materials are highly dependent on their constituent elements. As a typical representative of novel ternary layered MAX phase ceramic materials, Ta2AlC has attracted considerable attention from researchers due to its unique elemental composition, its potential performance advantages, and the current insufficient understanding of its synthesis–property correlation mechanisms and associated research demands. Therefore, extensive explorations and studies have been conducted on its synthesis techniques and performance characterization, with the aim of further uncovering and optimizing its potential applications. For instance, Jeitschko [18] was the first to discover the ternary layered ceramic Ta2AlC. To date, the experimental preparation methods for Ta2AlC materials include hot pressing (HP) and self-propagating high-temperature synthesis (SHS). Each of these methods possesses distinct advantages that render the fabrication of Ta2AlC materials feasible. Yeh et al. [19] and Tian et al. [20] employed SHS to investigate the influence of Al content on the synthesis of Ta2AlC. Manoun et al. [21] and Hu et al. [22] synthesized Ta2AlC via static pressing and analyzed the synthesis and decomposition mechanisms of Ta2AlC through reaction pathway analysis. Following the successful synthesis of Ta2AlC, Hu et al. [22] conducted characterization tests on its fundamental physical properties. The test results revealed that the compressive strength of Ta2AlC reached 804 MPa, a figure that underscores its remarkable performance as a damage-resistant material. Regarding electrical conductivity, as the temperature increased from 10 K to 300 K, the electrical conductivity of Ta2AlC correspondingly decreased from 37.8 × 106 W−1m−1 to 3.89 × 106 W−1m−1. Notably, when the temperature exceeded 70 K, the resistivity exhibited linear growth, indicating that Ta2AlC displayed typical metallic behavior. Additionally, within the temperature range from room temperature to 1100 °C, the thermal expansion coefficient of the material was 8.0 × 10−6 K−1, and from room temperature to 1227 °C, the thermal conductivity gradually decreased from 28.4 Wm−1K−1 to 25.5 Wm−1K−1. These data collectively confirm that Ta2AlC is an excellent conductor of heat and electricity and possesses superior mechanical properties.
Moreover, theoretical calculations have also played a significant role in the study of the ternary layered ceramic material Ta2AlC. For instance, Lin et al. [23] used density functional theory (DFT) to calculate the lattice parameters of Ta2AlC, which were in agreement with those measured by Jeitschko [13]. Qian et al. [24] employed first-principles calculations based on DFT to investigate the microscopic structure and electronic properties of Ta2AlC. Sun et al. [25] calculated the theoretical density of Ta2AlC to be 11.52 g/cm3 and its elastic modulus to be 318.6 GPa using first-principles methods. Music et al. [26] used ab initio calculations to study the shear behavior of Ta2AlC, and the results indicated that the C44 value is independent of the corresponding TaC, with the shear-resistant bands being filled by Ta atoms. Guo et al. [27] utilized DFT to explore the effects of vacancy defects on the structural, electronic, elastic, thermal conductivity, and optical properties of Ta2AlC. Lv et al. [28] conducted first-principles calculations on the electronic structure and optical properties of Ta2AlC, and the results suggested that Ta2AlC can be used as a temperature control coating material and an anti-ultraviolet coating material, with potential applications in the fields of aerospace and defense technologies.
In summary, research on Ta2AlC has primarily focused on its experimental synthesis and physical property testing, as well as theoretical calculations of its bulk microscopic structure, electronic properties, and optical characteristics. Doping is a crucial method for modulating the electronic structure and optical properties of materials. Therefore, theoretical studies on doping as a means of modifying the ternary layered ceramic material Ta2AlC hold noteworthy scientific guiding significance. However, studies on the doping modification of Ta2AlC are relatively rare. Thus, this paper takes Ta2AlC as the matrix and selects the elements V, Ga, and Si as dopants. Utilizing the first-principles pseudopotential plane-wave method, the study investigates the mechanical properties, electronic structure, and optical properties of doped Ta2AlC. It preliminarily reveals the physical mechanisms by which doping modulates the mechanical and photoelectric properties of Ta2AlC, providing a theoretical reference for experimental research on Ta2AlC materials.

2. Model and Method

The crystal structure of Ta2AlC belongs to the hexagonal crystal system with the space group P63/mmc (#194), and the lattice constants are a = b = 0.3079 nm and c = 1.3854 nm [29]. The coordinates of the non-equivalent atoms are as follows: Ta (1/3, 2/3, 1/12), Al (1/3, 2/3, 3/4), and C (0.0, 0.0, 0.0). The unit cell contains a total of 8 atoms, comprising 4 Ta atoms, 2 Al atoms, and 2 C atoms. In this study, a 2 × 2 × 2 supercell of Ta2AlC was selected for calculations, consisting of 64 atoms in total. Based on the 2 × 2 × 2 Ta2AlC supercell, substitutional doping was performed using two neighboring atoms from the same group as Ta-Al-C (such as V, Ga, and Si). The specific substitution methods were as follows: V replaced Ta, Ga replaced Al, and Si replaced C. The substitutional doping models of Ta2AlC selected for the calculations are shown in Figure 1.
The computational method employed in this study was the first-principles pseudopotential plane-wave method, with the primary calculations carried out using the Castep 8.0 software package [30]. The interaction between the ionic cores and electrons was treated using the ultrasoft pseudopotential (USPP) [31], while the exchange-correlation energy among electrons was handled using the Perdew–Burke–Ernzerhof (PBE) functional [32] within the generalized gradient approximation (GGA). The plane-wave cutoff energy was set at 300 eV, and the energy convergence criterion was 1.0 × 10⁻5 eV/atom. The valence electron configurations selected for the calculations were Ta 5d36s2, Al 3s23p1, C 2s22p2, V 3s23p63d34s2, Ga 3d104s24p1, and Si 3s23p2, with the remaining electrons treated as core electrons. The Brillouin zone integration adopted the symmetrical special k-point method in the form of a 5 × 5 × 1 Monkhorst-Pack [33] and set 32 × 32 × 144 FFT grid parameters.
The binding energy (Eb) is given by the expression [34]:
Eb = [Etot (TaxA1yCzMm) − xE(Ta) − yE(Al) − zE(C) − mE(M)]/(x + y + z + m)
where Etot (TaxA1yCzMm) is the total energy of the doped system, x is the number of Ta atoms in the system, y is the number of Al atoms in the system, z is the number of C atoms in the system, m is the number of dopant atoms M (such as V, Ga, or Si), E(Ta) is the total energy of an isolated Ta atom, E(Al) is the total energy of an isolated Al atom, E(C) is the total energy of an isolated C atom, and E(M) is the total energy of an isolated dopant atom M.

3. Results and Discussion

3.1. Geometric Structure

Changes in the elements and their positions at specific lattice sites within a crystal structure inevitably lead to systematic variations in interatomic bonding patterns. These changes can provide valuable guidance for the optimization and design of material properties. Therefore, the phonon spectrum, structural parameters, and binding energy of Ta2A1C before and after V, Ga, and Si doping are calculated in this paper, as shown in Figure 2 and Table 1.
We performed phonon spectrum calculations to evaluate the thermal stability of doped Ta2AlC. As shown in Figure 2, the phonon spectrum analysis reveals no imaginary frequencies, indicating that both undoped and doped Ta2AlC are thermally stable.
From Table 1, it can be seen that the lattice constants a and b of the 2 × 2 × 2 Ta2AlC supercell, after structural optimization, are within a 0.2% error margin of the results reported in the literature [27] (0.6320 nm), and the c/a ratio is within a 1.6% error margin of that reported in the literature [29]. These results indicate that the computational model used in this study can accurately describe the bonding state of Ta2AlC. Upon doping with V, Ga, and Si, the structural parameters of Ta2AlC undergo significant changes. The changes in lattice parameters are attributed to the differences in covalent radii of the substituted atoms. For instance, V (0.122 nm) replaces Ta (0.134 nm), Ga (0.126 nm) replaces Al (0.118 nm), and Si (0.111 nm) replaces C (0.077 nm). The data in Table 1 show that the trends in structural parameters of Ta2AlC after V and Si doping are consistent with the order of atomic sizes. However, after Ga doping, while the trends in lattice constants a and b are consistent with the order of their atomic sizes, the trends in lattice constant c and volume V do not match the order of their atomic sizes. This discrepancy may be due to the fact that along the c-axis direction, whether Ga and Ta can form bonds has a greater impact on the volume than the influence of atomic radii. Lattice defects or structural irregularities due to doping may result in significant changes in the mechanical and optical properties studied below.
The binding energies (Eb) listed in Table 1 for the doped Ta2AlC systems are all negative, indicating that the structures of all three doped systems are stable. The trend in formation energies is Eb(Ga) < Eb(V) < Eb(Si). From a thermodynamic perspective, the negative values of Eb suggest that the doped systems can be formed, and the Ga-doped system is more likely to form compared to the V- and Si doped systems.

3.2. Mechanical Properties

In this study, the elastic constants of Ta2AlC were calculated both before and after doping, and the effects of V, Ga, and Si doping on the mechanical properties of Ta2AlC were investigated. Table 2 lists the elastic constants of doped Ta2AlC.
For the Ta2AlC crystal with a hexagonal structure, according to the mechanical stability criterion, Born’s criterion [35], the following four conditions (2)–(5) must be simultaneously satisfied.
C11 > C12
(C11 + 2C12)C33 > 2C132
C44 > 0
C66 > 0
By processing the elastic constants listed in Table 2, it can be determined that the elastic constants of Ta2AlC before and after doping with V, Ga, and Si meet the aforementioned four conditions. This indicates that the systems studied in this paper possess mechanical stability. Additionally, it can be observed from Table 2 that, apart from a slight increase in C12 after V doping and a significant increase in C66 after Ga doping, the remaining elastic constants have decreased.
Based on the Voigt–Reuss–Hill method (VRH) [36,37,38], the elastic constants Cij listed in Table 2 can be used to derive the bulk modulus B, shear modulus G, Young’s modulus E, and Poisson’s ratio υ using the following Formulas (6)–(15). The results are presented in Table 3.
BV = (1/9)[2(C11 + C12) + C33 + 4C13]
GV = (1/30)(M + 3C11 − 3C12 + 12C44 + 6C66)
M = C11 + C12 + 2C33 − 4C13
C2 = (C11 + C12)C33 − 2C132
BR = C2/M
GR = 15{(18BV/C2) + [6/(C11 − C12)] + (6/C44) + (3/C66)}−1
B = (BV + BR)/2
G = (GV + GR)/2
E = 9BG/(3B + G)
υ = (3B − 2G)/[2(3B + G)]
Table 3 lists the mechanical properties parameters of doped Ta2AlC. As can be seen from Table 3, the Young’s modulus of Ta2AlC is approximately 317 GPa, which deviates by no more than 2.5% from the result reported by Guo [27] (325 GPa) is larger than that of Ti2A1C (305 GPa) of the same material type. Due to its high Young’s modulus and shear modulus, Ta2AlC exhibits exceptional rigidity and resistance to deformation, making it a promising candidate material for manufacturing aerospace engine components and aircraft structural parts. After doping with V and Si, the bulk modulus B, shear modulus G, and Young’s modulus E of Ta2AlC all decrease, while the Poisson’s ratio υ and the B/G ratio increase. This indicates that the material’s ability to resist deformation is weakened while its plasticity is enhanced after doping with V and Si. In contrast to the situation after doping with V and Si, the Poisson’s ratio υ and the B/G ratio of Ta2AlC doped with Ga decrease, indicating that the introduction of Ga strengthens the material’s ability to resist lateral deformation and increases its brittleness.
The bond energy plays a regulatory role in lattice distortion, and this in turn significantly affects the plasticity and brittleness of materials. Due to differences in bond energy between the substituted atoms and the original atoms—for instance, V (514 kJ/mol) replacing Ta (782 kJ/mol), Ga (271 kJ/mol) replacing Al (330 kJ/mol), and Si (326 kJ/mol) replacing C (346 kJ/mol)—the plasticity or brittleness of the doped material undergoes changes. When V and Si are introduced as dopants, the lower bond energy of the substituting atoms weakens their ability to suppress lattice distortion caused by atomic size mismatch. This facilitates dislocation motion within the lattice, thereby enhancing the plasticity of Ta2AlC. Conversely, Ga doping hinders the generation of dislocations in the lattice, leading to the increased brittleness of Ta2AlC. This is also consistent with the significant increase in C66 after Ga doping shown in Table 2. The main reason for brittleness is the introduction of Ga; whether Ga and Ta can form bonds along the c-axis becomes the main consideration in exploring brittleness. This is the same reason for the change in lattice parameters analyzed earlier.
The arrangement of atoms in a crystal and the different environments in various directions lead to different physical properties in different directions, thus giving rise to anisotropy. Elastic anisotropy is represented by the elastic anisotropy index, which is specifically characterized by the general elastic anisotropy index AU, the compressional anisotropy percentage Acomp, and the shear anisotropy percentage Ashear. In addition, the shear anisotropy factors (A1, A2, and A3) can be used to indicate the degree of anisotropy of the material on the (001), (010), and (100) planes, respectively. Therefore, based on the data in Table 2 and Table 3, the elastic anisotropy indices and shear anisotropy factors of Ta2AlC before and after doping were calculated. The specific calculation formulas [39,40,41,42] are shown in Equations (16)–(21). The results of the elastic anisotropy calculations are presented in Table 4.
AU = 5GV/GR + BV/BR − 6
Acomp = (BV − BR)/(GV + GR) × 100%
Ashear = (GV − GR)/(GV + GR) × 100%
A1 = 4C44/(C11 − 2C13 + C33)
A2 = 4C44/(C11 − 2C13 + C33)
A3 = 4C66/(C11 − 2C13 + C33)
AU can evaluate the anisotropy of material. If the material is isotropic, AU = 0. The deviation of AU from zero indicates the degree of anisotropy of the material. As can be seen from Table 4, before doping, AU = 0.2307, indicating that Ta2A1C is anisotropic, which is very consistent with the result (0.237) in reference [27]. When doped with V, Ga, or Si, the AU, Ashear, A1, A2, and A3 values of the Ta2A1C material increase, indicating that the degree of anisotropy of the material increases. The increase in anisotropy is due to lattice distortion caused by doping. According to anisotropy, the cause is that the directional difference of the atomic ar-rangement causes the physical properties to change with the direction. Due to the incorporation of V, Ga, or Si dopants, the position of each atom in the cell is shifted, which leads to changes in the lattice parameters of Ta2AlC, so that the lattice is distorted, and these changes directly affect the change in the anisotropy index, manifesting in the enhanced anisotropy of the Ta2AlC material. However, it is worth noting that after Si doping, Acomp = 0, indicating that the compression anisotropy of Ta2A1C can be effectively eliminated after Si replaces the C atom through doping. However, the fact that Ashear = 3.98% indicates that shear anisotropy still exists and trends upward, so the material is still anisotropic on the whole.

3.3. Electronic Structure

Figure 3 shows the band structure of Ta2A1C before and after V, Ga, and Si doping at energies ranging from −14 eV to 2 eV. Figure 3 shows the band structure in the Brillouin region along the direction of highly symmetric points, and it can be seen that Ta2A1C does not have a band gap. It is shown that the valence band and conduction band near the Fermi level (EF = 0 eV) overlap each other; that is, the Fermi level passes through multiple energy bands of A-H, K-Γ, Γ-F, and Q-H. It can be seen that Ta2AlC is a conductor with excellent conductivity, and its conductive characteristics are not weakened after doping with V, Ga, or Si. After doping, the energy band increases significantly, which indicates that the introduction of impurity atoms changes the interaction between different atoms.
Figure 4 shows the total electronic density of states of Ta2AlC before and after doping. As can be seen from Figure 4, the electronic density of states is non-zero near the Fermi level, indicating that Ta2AlC is a conductor, which is consistent with the previous analysis of the band structure. In the density of states diagrams, the electronic density of states for spin-up and spin-down electrons is symmetrically distributed, indicating that the Ta2AlC system is non-magnetic both before and after doping.
To elucidate the mechanisms by which dopant atoms influence the band structure and electronic density of states, partial density of states diagrams for each atom in Ta2AlC before and after doping are specifically plotted, as shown in Figure 5, Figure 6, Figure 7 and Figure 8. As can be seen from Figure 5, the electronic density of states between −14 eV and −10 eV is primarily contributed by the 2s states of C and the 5d and 6s states of Ta, with an electronic transition peak existing between Ta-5d and C-2s at the valence band −11.71 eV. The electrons between −10 eV and −3.41 eV are mainly contributed to by the 2p states of C and the 5d states of Ta, with an electronic transition peak existing between Ta-5d and C-2p at the valence band −4.82 eV. The electrons between −3.41 eV and 2.50 eV are predominantly contributed to by the 5d states of Ta and the 3p states of Al, with electronic transition peaks existing between Ta-5d and Al-3p at the valence band −1.87 eV and the conduction band 1.06 eV.
As can be seen from Figure 6, after V doping, two new peaks appear in the electronic density of states between −64.54 eV and −38.29 eV, which are attributed to the contributions of the V-3p and V-4s states. The emergence of these new peaks is consistent with the changes in the band structure shown in Figure 3 (the two new bands are not depicted in the figure). The introduction of a very small number of V-3d states weakens the C-2s states between −14 eV and −10 eV, resulting in a decrease in the electronic transition peak between Ta-5d and C-2s. This corresponds to the sparser bands in Figure 3 and the reduced density of states peak in Figure 4. Additionally, near the Fermi level, the introduction of V-3d states increases the number and density of bands in this region, leading to a slight increase in the total electronic density of states, which is consistent with the results shown in Figure 3 and Figure 4.
As depicted in Figure 7, after Ga doping, a new peak in the electronic density of states emerges near −14.67 eV, attributed to the contribution of Ga-4d states. The appearance of this new peak aligns with the changes in the band structure illustrated in Figure 3 (the new band is not depicted in the figure). Near −2.57 eV, the introduction of Ga-4p states leads to electronic transitions between Ga-4p and Ta-5d, resulting in an increase in the number and density of bands in this region, as well as an enhancement in the density of states peak. This indicates that the introduction of Ga strengthens the covalent interaction between Ta and Al, which is consistent with the analysis of lattice parameters and mechanical properties parameters presented in Table 1, Table 2 and Table 3.
As depicted in Figure 8, after Si doping, near −12.07 eV, the introduction of Si-3s states weakens the C-2s states, resulting in a reduction in the electronic transition peak between Ta-5d and C-2s. This corresponds to the sparser bands in Figure 3 and the reduced density of states peak in Figure 4. Near −9.60 eV and −8.62 eV, the introduction of Si-3s states results in the non-zero total electronic density of states in Figure 4, corresponding to the appearance of two new bands in Figure 3. Additionally, near the Fermi level, the introduction of Si-3p states leads to electronic transitions between Si-3p and Ta-5d, resulting in an increase in the number and density of bands in this region, as well as an enhancement in the density of states peak. This indicates a strong covalent interaction between Si and Ta. In conjunction with the supercell model of Si doped Ta2AlC shown in Figure 2, the replacement of C atoms with Si atoms at the four crystal edges and the center of the unit cell effectively eliminates the compressive anisotropy of Ta2AlC.

3.4. Mulliken Population Analysis

Table 5 presents the Mulliken population analysis of adjacent bonds between atoms in doped Ta2AlC. As indicated in Table 5, in undoped Ta2AlC, the population value of the C-Ta bond is 0.44, while that of the Al-Ta bond is 0.36, indicating that the covalent bond between C and Ta is stronger than that between Al and Ta.
After the V doping of Ta2AlC, the population value of the C-Ta bond increases to 0.52, and that of the Al-Ta bond increases to 0.39, suggesting that the introduction of V enhances the covalent interactions between C-Ta and Al-Ta atoms. Additionally, V atoms form chemical bonds with neighboring C and Al atoms, with population values of 0.23 and 0.22, respectively. The above results indicate that the introduction of V enhances the covalent interaction between atoms, which clearly explains the reason for the enhanced plasticity of the Ta2A1C material after doping with V reported in Section 3.2.
Following Ga doping in Ta2AlC, the population value of the C-Ta bond rises to 0.46, and that of the Al-Ta bond rises to 0.39, indicating that Ga doping strengthens the covalent interactions between C-Ta and Al-Ta atoms. However, the population value of the Ga-Ta bond between Ga atoms and neighboring Ta atoms is −0.64, suggesting that no chemical bond is formed between Ga and Ta. Combined with V doped Ta2AlC models, as shown in Figure 2c, along the c-axis direction, the interaction between the two layers of Ta atoms is significantly weakened because Ga-Ta does not form a chemical bond, which clearly explains the reason why the Ta2A1C material becomes brittle after Ga doping, as reported in Section 3.2.
After the Si doping of Ta2AlC, the population value of the C-Ta bond decreases to 0.43, and that of the Al-Ta bond decreases to 0.33, indicating that Si doping weakens the covalent interactions between C-Ta and Al-Ta atoms. Meanwhile, Si atoms form chemical bonds with neighboring Ta atoms, attaining a population value of 0.50, indicating a strong covalent interaction between Si and Ta. Unlike V and Ga doping, Ta-Ta bonds are formed after Si doping, with a population value of 0.12. Due to the strong covalent interaction between Si and Ta atoms and the bonding between Ta atoms, the plasticity of Ta2A1C doped with Si becomes stronger.

3.5. Optical Properties

To investigate the impact of doping on the optical properties of Ta2AlC, this study calculated the optical properties of this material, including the complex dielectric function, absorption coefficient, reflectivity, refractive index, and loss function.
The complex dielectric function is of paramount importance for studying the optical properties of metallic materials. Within the linear response regime, the complex dielectric function ε(ω) can be expressed as:
ε(ω) = ε1(ω) + ε2(ω)
Here, ε1(ω) represents the real part of the dielectric function, while ε2(ω) represents the imaginary part.
Figure 9 depicts the variation in the real and imaginary parts of the dielectric function of doped Ta2AlC with energy.
As can be seen from Figure 9a, the static dielectric constant of Ta2AlC, ε1(0), is 79.63. After doping with V and Si, the static dielectric constant increases significantly to 124.18 (V) and 128.32 (Si), respectively, while doping with Ga results in a decrease to 76.14. These changes in the static dielectric constant suggest potential applications for optical components where specific dielectric constants are required. Furthermore, it can be observed that within the energy range of E < 0.55 eV (in the infrared region), the real part of the dielectric function ε1(ω) is much larger for Ta2AlC doped with V and Si than that of the undoped material. This indicates a significant enhancement in the ability of V- and Si doped Ta2AlC to store electromagnetic energy. In contrast, Ga doping results in almost no change in ε1(ω).
The imaginary part of the dielectric function ε2(ω) serves as a bridge between the microscopic physical processes of interband transitions and the electronic structure of solids, reflecting the electronic structure of solids and various spectral information. As can be seen from Figure 9b, Ta2AlC has a single dielectric peak in ε2(ω) with a value of 24.55, located near E = 0.38 eV. After V doping, the dielectric peak significantly increases to 41.88 and exhibits a slight red shift towards lower energies. Following Ga doping, the dielectric peak slightly decreases to 24.18 and shows a noticeable blue shift towards higher energies. After Si doping, the position of the dielectric peak remains almost unchanged, but the peak value significantly increases to 42.26. Within the energy range of E < 1.58 eV (in the infrared region), the imaginary part of the dielectric function ε2(ω) is much larger for Ta2AlC doped with V and Si than that of the undoped material, indicating a significant enhancement in dielectric loss after V and Si doping. In conjunction with the partial density of states diagrams, the changes in the dielectric peaks can be attributed to the hybridization of V-3d, Ga-4p, and Si-3p states with Ta-5d and Al-3p states near the Fermi level, which alters the electronic transitions within and between bands, thereby affecting the dielectric function and other optical properties of Ta2AlC.
Figure 10 depicts the optical properties of doped Ta2AlC. From Figure 10a, it can be observed that when E = 0 eV, the absorption coefficients of doped Ta2AlC materials are all non-zero. As the energy E increases, the absorption coefficient rises sharply, reaching a peak near E = 5.01 eV (ultraviolet region). This indicates that the doped Ta2AlC can absorb photons ranging from low-frequency electromagnetic waves to those in the ultraviolet region, implying the absence of an optical band gap. These results are consistent with those reported in refs. [27,28]. Notably, after V doping, a new absorption peak appears near E = 39.29 eV, attributed to electronic transitions from the V-3p state to the Fermi level. As can be seen from Figure 10b, the reflectivity of Ta2AlC achieves its maximum value of 95% near E = 9.53 eV (in the ultraviolet region). The exceptionally high reflectivity of Ta2AlC (up to 95%) primarily stems from free electron effects, driven by its high carrier concentration, which induces a metal-like plasmonic response [43]. Specifically, the metallic bonding (Ta-Al bonds) in Ta2AlC provides a substantial number of free electrons, forming a metal-like electron gas. Under the oscillating electric field of incident light, these free electrons undergo plasmon oscillations, resulting in high reflectivity for photons with energies below the plasma frequency. After doping with V, Ga, and Si, the reflectivity peak slightly decreases to 94%, 94%, and 92%, respectively. These results indicate that both doped and undoped Ta2AlC can serve as potential anti-ultraviolet coating materials. Moreover, within the energy range of E < 8.66 eV, Si doping significantly enhances the reflectivity of Ta2AlC. It can also be observed that in the visible light region between 1.58 eV and 3.18 eV, the photon frequency has minimal impact on the reflectivity, which remains around 50% for both doped and undoped Ta2AlC. This suggests that Ta2AlC, before and after doping, exhibits a similar metallic gray appearance, indicating its potential as a temperature control coating material in the visible light region. As can be seen from Figure 10c, the refractive index of Ta2AlC is n0 = 8.99, which changes to n0(V) = 11.28, n0(Ga) = 8.79, and n0(Si) = 11.43 after doping with V, Ga, and Si, respectively. For optical devices made from Ta2AlC material, where specific refractive index values are required, these results can be used as a reference. As can be seen from Figure 10d, the loss function of Ta2AlC reaches its peak value of 247.82 near E = 9.75eV (in the ultraviolet region). After doping with V, Ga, and Si, the peak of the loss function is significantly reduced. However, the light loss of Ta2AlC is confined to this energy range, indicating that both doped and undoped Ta2AlC materials can be used as light storage materials.

4. Conclusions

In this study, the geometric structure, mechanical properties, electronic structure, Mulliken population analysis, and optical properties of Ta2AlC doped with V, Ga, and Si were investigated using the first-principles pseudopotential plane-wave method. The results show the following:
(1)
The structural parameters of Ta2AlC undergo significant changes after doping with V, Ga, and Si. The variations in lattice parameters are attributed to the differences in covalent radii of the substituted atoms. The mechanical stability of bulk Ta2AlC and Ta2AlC doped with V, Ga, and Si were confirmed using Born’s criterion.
(2)
The mechanical properties analysis of doped Ta2AlC reveal that V and Si doping weaken the material’s resistance to deformation while enhancing its plasticity. In contrast, Ga doping increases the material’s resistance to lateral deformation and brittleness. The elastic anisotropy indices (AU, Ashear, A1, A2, and A3) of Ta2AlC increase after V, Ga, and Si doping, indicating an enhanced degree of anisotropy.
(3)
The electronic structure calculations demonstrate that Ta2AlC is a conductor with excellent electrical conductivity, which is not diminished by V, Ga, or Si doping. The symmetric distribution of spin-up and spin-down electronic states density confirms that the Ta2AlC system remains non-magnetic after doping. The partial density of states diagrams successfully elucidate the mechanisms by which dopant atoms influence the band structure and electronic states density, providing results consistent with the analysis of the material’s lattice parameters and mechanical properties.
(4)
The Mulliken population analysis indicates that V and Ga doping enhance the covalent interactions between C-Ta and Al-Ta atoms, while Si doping weakens these interactions.
(5)
The optical properties calculations reveal that V and Si doping significantly enhance the electromagnetic energy storage capacity and dielectric loss of Ta2AlC, whereas Ga doping has almost no effect. The reflectivity of Ta2AlC in the ultraviolet region peaks at 95% for the undoped material and decreases to 94%, 94%, and 92% after V, Ga, and Si doping, respectively, indicating that both doped and undoped Ta2AlC can serve as potential anti-ultraviolet coating materials. In the visible light region, doped and undoped Ta2AlC exhibit a similar metallic gray appearance, suggesting their potential as temperature control coating materials in the visible light region. The light loss of Ta2AlC is confined to a narrow energy range, unaffected by doping, indicating its suitability as a light storage material.
The above results demonstrate that different dopants can effectively modulate the mechanical and photoelectric properties of Ta2AlC.

Author Contributions

Data curation, Z.Z. and X.Q.; formal analysis, Z.Z.; funding acquisition, W.Y.; methodology, C.Z.; writing—original draft, Z.Z., X.Q. and W.Y.; writing—review and editing, C.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Key Laboratory of Materials Simulation and Computing of Anshun University (Grant no. Asxyxkpt201803) and the Project of the Education Department of Guizhou Province: Guizhou Provincial University Integrated key tackling platform (Grant no. [2021]315).

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Acknowledgments

The authors acknowledge the High Performance Computing Center of Anshun University of China for the support.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Barsoum, M.W. The Mn+1AXn Phases: A New Class of Solids; Thermodynamically Stable Nanolaminates. Prog. Solid State Chem. 2000, 28, 201–281. [Google Scholar] [CrossRef]
  2. Barsoum, M.W.; El-Raghy, T. The MAX Phases: Unique New Carbide and Nitride Materials. Am. Sci. 2001, 89, 334–343. [Google Scholar] [CrossRef]
  3. Ganguly, A.; Zhen, T.; Barsoum, M.W. Synthesis and Mechanical Properties of Ti3GeC2 and Ti3(SixGe1−x)C2 (x = 0.5, 0.75) Solid Solutions. J. Alloys Compd. 2004, 376, 287–295. [Google Scholar] [CrossRef]
  4. Hwang, S.S.; Lee, S.C.; Han, J.H.; Lee, D.; Park, S.-W. Machinability of Ti3SiC2 with Layered Structure Synthesized by Hot Pressing Mixture of TiCx and Si Powder. J. Eur. Ceram. Soc. 2012, 32, 3493–3500. [Google Scholar] [CrossRef]
  5. Barsoum, M.W.; Salama, I.; El-Raghy, T.; Golczewski, J.; Seifert, H.J.; Aldinger, F.; Porter, W.D.; Wang, H. Thermal and Electrical Properties of Nb2AlC, (Ti,Nb)2AlC and Ti2AlC. Metall. Mater. Trans. A 2002, 33, 2775–2779. [Google Scholar] [CrossRef]
  6. Wang, X.H.; Zhou, Y.C. Solid-Liquid Reaction Synthesis and Simultaneous Densification of Polycrystalline Ti2AlC. Z. Met. 2002, 93, 66–71. [Google Scholar] [CrossRef]
  7. Wang, X.H.; Zhou, Y.C. Intermediate-Temperature Oxidation Behavior of Ti2AlC in Air. J. Mater. Res. 2002, 17, 2974–2981. [Google Scholar] [CrossRef]
  8. Wang, X.H.; Zhou, Y.C. High-Temperature Oxidation Behavior of Ti2AlC in Air. Oxid. Met. 2003, 59, 303–320. [Google Scholar] [CrossRef]
  9. Sundberg, M.; Malmqvist, G.; Magnusson, A.; El-Raghy, T. Alumina Forming High Temperature Silicides and Carbides. Ceram. Int. 2004, 30, 1899–1904. [Google Scholar] [CrossRef]
  10. Bouhemadou, A.; Khenata, R. Prediction Study of Structural and Elastic Properties under the Pressure Effect of M2GaC (M = Ti, V, Nb, Ta). J. Appl. Phys. 2007, 102, 043528. [Google Scholar]
  11. Ying, G.B.; He, X.D.; Li, M.W.; Han, W.; He, F.; Du, S. Synthesis and Mechanical Properties of High-Purity Cr2AlC Ceramic. Mater. Sci. Eng. A 2011, 528, 2635–2640. [Google Scholar] [CrossRef]
  12. Barsoum, M.W.; El-Raghy, T.; Ogbuji, L.U.J.T. Oxidation of Ti3SiC2 in Air. J. Electrochem. Soc. 1997, 144, 2508–2516. [Google Scholar] [CrossRef]
  13. Su, X.J.; Liu, G.X.; Dai, B.; Wan, D.; Bao, Y.; Feng, Q.; Grasso, S.; Hu, C. Ablation Mechanisms of Ti3SiC2 Ceramic at 1600 °C in Nitrogen Plasma Flame. Ceram. Int. 2022, 48, 14004–14013. [Google Scholar]
  14. Ding, W.; Hu, B.T.; Fu, S.; Wan, D.; Bao, Y.; Feng, Q.; Grasso, S.; Hu, C. Ultra-Fast Thermal Shock Evaluation of Ti2AlC Ceramic. Materials 2022, 15, 6877. [Google Scholar] [CrossRef]
  15. Su, X.J.; Hu, B.T.; Quan, Y.; Qin, Y.; Feng, Q.; Hu, C. Ablation Behavior and Mechanism of Bulk MoAlB Ceramic at 1670–2550 °C in Air Plasma Flame. J. Eur. Ceram. Soc. 2021, 41, 5474–5483. [Google Scholar]
  16. Gonzalez-Julian, J. Processing of MAX Phases: From Synthesis to Applications. J. Am. Ceram. Soc. 2021, 104, 659–690. [Google Scholar]
  17. Zhang, Q.Q.; Zhou, Y.C.; San, X.Y.; Li, W.; Bao, Y.; Feng, Q.; Grasso, S.; Hu, C. Zr2SeB and Hf2SeB: Two New MAB Phase Compounds with the Cr2AlC-Type MAX Phase (211 Phase) Crystal Structures. J. Adv. Ceram. 2022, 11, 1764–1776. [Google Scholar] [CrossRef]
  18. Jeitschko, W.; Nowotny, H.; Benesovsky, F. Carbon-Containing Ternary Compounds (H-Phase). Monatsh. Chem. 1963, 94, 672–676. [Google Scholar]
  19. Yeh, C.L.; Shen, Y.G. Effects of Al Content on Formation of Ta2AlC by Self-Propagating High-Temperature Synthesis. J. Alloys Compd. 2009, 482, 219–223. [Google Scholar] [CrossRef]
  20. Tian, B.N.; Ying, G.B.; Wang, P.J.; Wang, C.; Wu, Y.P. In-Situ Synthesis Mechanism of Nano-Layered Ternary Ta2AlC. J. Synth. Cryst. 2015, 44, 1773–1777. [Google Scholar]
  21. Manoun, B.; Gulve, R.P.; Saxena, S.K.; Gupta, S.; Barsoum, M.W.; Zha, C.S. Compression Behavior of M2AlC (M = Ti, V, Cr, Nb, and Ta) Phases to above 50 GPa. Phys. Rev. B 2006, 73, 024110. [Google Scholar] [CrossRef]
  22. Hu, C.F.; Zhang, J.; Bao, Y.W.; Wang, J.; Li, M.; Zhou, Y. In-Situ Reaction Synthesis and Decomposition of Ta2AlC. Int. J. Mater. Res. 2008, 99, 8–13. [Google Scholar] [CrossRef]
  23. Lin, Z.J.; Zhuo, M.J.; Zhou, Y.C.; Li, M.; Wang, J. Microstructures and Theoretical Bulk Modulus of Layered Ternary Tantalum Aluminum Carbides. J. Am. Ceram. Soc. 2006, 89, 3765–3769. [Google Scholar]
  24. Qian, X.K.; Li, Y.B.; He, X.D.; Fan, H.; Yun, S. First-Principles Study of Structural and Electronic Properties of Ternary Layered Ta2AlC. J. Phys. Chem. Solids 2011, 72, 954–956. [Google Scholar] [CrossRef]
  25. Sun, Z.M.; Li, S.; Ahuja, R.; Schneider, J.M. Calculated Elastic Properties of M2AlC (M = Ti, V, Cr, Nb and Ta). Solid State Commun. 2004, 129, 589–592. [Google Scholar] [CrossRef]
  26. Music, D.; Sun, Z.M.; Voevodin, A.A.; Schneider, J.M. Electronic Structure and Shearing in Nanolaminated Ternary Carbides. Solid State Commun. 2006, 139, 139–143. [Google Scholar] [CrossRef]
  27. Guo, Z.Z. First-Principles Studies of the Effects of Vacancy Defects on the Properties of Ta2AlC. Phys. B Condens. Matter 2023, 668, 415160. [Google Scholar] [CrossRef]
  28. Lv, L.X.; Xu, W.W.; Zhu, C.C. First-Principles Study on the Electronic Structure and Optical Properties of Ta2AlC. J. Synth. Cryst. 2016, 45, 520–524. [Google Scholar]
  29. Hu, C.F.; He, L.F.; Zhang, J.; Bao, Y.; Wang, J.; Li, M.; Zhou, Y. Microstructure and Properties of Bulk Ta2AlC Ceramic Synthesized by an In Situ Reaction/Hot Pressing Method. J. Eur. Ceram. Soc. 2008, 28, 1679–1685. [Google Scholar] [CrossRef]
  30. Segall, M.D.; Lindan, P.J.D.; Probert, M.J.; Pickard, C.J.; Hasnip, P.J.; Clark, S.J.; Payne, M.C. First-Principles Simulation: Ideas, Illustrations and the CASTEP Code. J. Phys. Condens. Matter 2002, 14, 2717–2744. [Google Scholar] [CrossRef]
  31. Vanderbilt, D. Soft Self-Consistent Pseudopotentials in a Generalized Eigenvalue Formalism. Phys. Rev. B 1990, 41, 7892–7895. [Google Scholar] [CrossRef]
  32. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [PubMed]
  33. Monkhorst, H.J.; Pack, J.D. Special Points for Brillouin-Zone Integrations. Phys. Rev. B 1976, 13, 5188–5192. [Google Scholar] [CrossRef]
  34. Xiao, B.; Feng, J.; Zhou, C.T.; Xing, J.; Xie, X.; Chen, Y. First Principles Study on the Electronic Structures and Stability of Cr7C3 Type Multi-Component Carbides. Chem. Phys. Lett. 2008, 459, 129–132. [Google Scholar] [CrossRef]
  35. Watt, J.P.; Peselnick, L. Clarification of the Hashin-Shtrikman Bounds on the Effective Elastic Moduli of Polycrystals with Hexagonal, Trigonal, and Tetragonal Symmetries. J. Appl. Phys. 1980, 51, 1525–1531. [Google Scholar] [CrossRef]
  36. Reuss, A. Berechnung der Fließgrenze von Mischkristallen auf Grund der Plastizitätsbedingung für Einkristalle. ZAMM 1929, 9, 49–58. [Google Scholar] [CrossRef]
  37. Hill, R. The Elastic Behaviour of a Crystalline Aggregate. Proc. Phys. Soc. A 1952, 65, 349–354. [Google Scholar] [CrossRef]
  38. Hadi, M.A.; Kelaidis, N.; Naqib, S.H. Mechanical Behaviors, Lattice Thermal Conductivity and Vibrational Properties of a New MAX Phase Lu2SnC. J. Phys. Chem. Solids 2019, 129, 162–171. [Google Scholar] [CrossRef]
  39. Chung, D.H.; Buessem, W.R.; Vahldiek, F.W.; Mersol, S.A. Anisotropy in Single Crystal Refractory Compounds; Plenum Press: New York, NY, USA, 1968. [Google Scholar]
  40. Huang, B.; Duan, Y.H.; Hu, W.C.; Sun, Y.; Chen, S. Structural, Anisotropic Elastic and Thermal Properties of MB (M = Ti, Zr and Hf) Monoborides. Ceram. Int. 2015, 41, 6831–6843. [Google Scholar] [CrossRef]
  41. Zou, M.T.; Bao, L.K.; Yang, A.C.; Duan, Y.; Peng, M.; Cao, Y.; Li, M. Structural, Elastic, Mechanical, Electronic, Damage Tolerance, Fracture Toughness, and Optical Properties of Cr-Al-B MAB Phases Studied by First-Principles Calculations. J. Mater. Res. 2023, 38, 1396–1409. [Google Scholar] [CrossRef]
  42. Li, P.; Ma, L.; Peng, M.; Shu, B.; Duan, Y. Elastic Anisotropies and Thermal Conductivities of WB2 Diborides in Different Crystal Structures: A First-Principles Calculation. J. Alloys Compd. 2018, 747, 905–915. [Google Scholar] [CrossRef]
  43. Maier, S.A. Plasmonics: Fundamentals and Applications; Springer: New York, NY, USA, 2007. [Google Scholar]
Figure 1. Doping models of the 2 × 2 × 2 Ta2AlC: (a) Ta2AlC, (b) V doped Ta2AlC, (c) Ga doped Ta2AlC, and (d) Si doped Ta2AlC.
Figure 1. Doping models of the 2 × 2 × 2 Ta2AlC: (a) Ta2AlC, (b) V doped Ta2AlC, (c) Ga doped Ta2AlC, and (d) Si doped Ta2AlC.
Crystals 15 00309 g001
Figure 2. Phonon spectrum of doped Ta2A1C. (a) Ta2AlC, (b) V doped Ta2AlC, (c) Ga doped Ta2AlC, and (d) Si doped Ta2AlC.
Figure 2. Phonon spectrum of doped Ta2A1C. (a) Ta2AlC, (b) V doped Ta2AlC, (c) Ga doped Ta2AlC, and (d) Si doped Ta2AlC.
Crystals 15 00309 g002aCrystals 15 00309 g002b
Figure 3. Band structure of doped Ta2A1C: (a) Ta2AlC, (b) V doped Ta2AlC, (c) Ga doped Ta2AlC, and (d) Si doped Ta2AlC.
Figure 3. Band structure of doped Ta2A1C: (a) Ta2AlC, (b) V doped Ta2AlC, (c) Ga doped Ta2AlC, and (d) Si doped Ta2AlC.
Crystals 15 00309 g003
Figure 4. Density of states of doped Ta2A1C: (a) Ta2AlC, (b) V doped Ta2AlC, (c) Ga doped Ta2AlC, and (d) Si doped Ta2AlC.
Figure 4. Density of states of doped Ta2A1C: (a) Ta2AlC, (b) V doped Ta2AlC, (c) Ga doped Ta2AlC, and (d) Si doped Ta2AlC.
Crystals 15 00309 g004aCrystals 15 00309 g004b
Figure 5. Partial density of states of Ta2A1C: (a) Ta atom, (b) Al atom, and (c) C atom.
Figure 5. Partial density of states of Ta2A1C: (a) Ta atom, (b) Al atom, and (c) C atom.
Crystals 15 00309 g005
Figure 6. Partial density of states of V doped Ta2A1C: (a) Ta atom, (b) Al atom, (c) C atom, and (d) V atom.
Figure 6. Partial density of states of V doped Ta2A1C: (a) Ta atom, (b) Al atom, (c) C atom, and (d) V atom.
Crystals 15 00309 g006
Figure 7. Partial density of states of Ga doped Ta2A1C: (a) Ta atom, (b) Al atom, (c) C atom, and (d) Ga atom.
Figure 7. Partial density of states of Ga doped Ta2A1C: (a) Ta atom, (b) Al atom, (c) C atom, and (d) Ga atom.
Crystals 15 00309 g007
Figure 8. Partial density of states of Si doped Ta2A1C: (a) Ta atom, (b) Al atom, (c) C atom, and (d) Si atom.
Figure 8. Partial density of states of Si doped Ta2A1C: (a) Ta atom, (b) Al atom, (c) C atom, and (d) Si atom.
Crystals 15 00309 g008
Figure 9. Complex dielectric function of doped Ta2AlC: (a) ε1(ω), (b) ε2(ω).
Figure 9. Complex dielectric function of doped Ta2AlC: (a) ε1(ω), (b) ε2(ω).
Crystals 15 00309 g009
Figure 10. Optical properties of doped Ta2AlC: (a) absorption coefficient, (b) reflectivity, (c) refractive index, and (d) loss function.
Figure 10. Optical properties of doped Ta2AlC: (a) absorption coefficient, (b) reflectivity, (c) refractive index, and (d) loss function.
Crystals 15 00309 g010
Table 1. Structural parameters and binding energies of doped Ta2AlC.
Table 1. Structural parameters and binding energies of doped Ta2AlC.
Samplea (nm)b (nm)c (nm)V (nm3)E (eV)Eb (eV)
Ta2A1C0.63090.63092.79410.9633−7797.3365−8.3016
V doped Ta2A1C0.62870.62872.78490.9534−11,475.9034−8.2683
Ga doped Ta2A1C0.63150.63152.78830.9631−11,789.8840−8.2742
Si doped Ta2A1C0.63770.63772.80500.9858−7696.9951−8.1058
Table 2. Elastic constants of doped Ta2AlC.
Table 2. Elastic constants of doped Ta2AlC.
SampleC11C12C13C33C44C66
Ta2A1C348.7114114.0028128.6486312.5064164.5408118.6462
V doped Ta2A1C341.0590114.4546126.3630299.7501163.4924115.7912
Ga doped Ta2A1C335.561494.7569118.5429281.2559164.4988121.4161
Si doped Ta2A1C299.708990.5178121.9969263.7427152.3105102.3365
Table 3. Mechanical properties parameters of doped Ta2AlC.
Table 3. Mechanical properties parameters of doped Ta2AlC.
SampleBGEυB/G
Ta2A1C194.6353129.2186317.41240.22821.5062
V doped Ta2A1C190.5174126.2935310.31210.22851.5085
Ga doped Ta2A1C179.3631127.7701309.75790.21221.4038
Si doped Ta2A1C170.2374112.1657275.90190.22991.5177
Table 4. Elastic anisotropy indices of doped Ta2AlC.
Table 4. Elastic anisotropy indices of doped Ta2AlC.
SampleAUAcompAshearA1A2A3
Ta2A1C0.2307 0.07%2.25%1.6294 1.6294 1.1749
V doped Ta2A1C0.2652 0.14%2.57%1.6851 1.6851 1.1935
Ga doped Ta2A1C0.2989 0.16%2.88%1.7328 1.7328 1.2790
Si doped Ta2A1C0.4150 0.00%3.98%1.9071 1.9071 1.2814
Table 5. Mulliken population analysis of doped Ta2AlC.
Table 5. Mulliken population analysis of doped Ta2AlC.
SampleBondPopulationLength (nm)
Ta2A1CC-Ta0.440.22
Al-Ta0.360.29
V doped Ta2A1CC-Ta0.520.22
Al-Ta0.390.28
V-C0.230.21
V-Al0.220.29
Ga doped Ta2A1CC-Ta0.460.22
Al-Ta0.390.29
Ga-Ta−0.640.27
Si doped Ta2A1CC-Ta0.430.23
Al-Ta0.330.28
Si-Ta0.500.24
Ta-Ta0.120.30
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

Zhang, Z.; Zhang, C.; Qin, X.; Yan, W. Research on the Mechanical and Photoelectric Properties Regulation of the New-Type Ceramic Material Ta2AlC. Crystals 2025, 15, 309. https://doi.org/10.3390/cryst15040309

AMA Style

Zhang Z, Zhang C, Qin X, Yan W. Research on the Mechanical and Photoelectric Properties Regulation of the New-Type Ceramic Material Ta2AlC. Crystals. 2025; 15(4):309. https://doi.org/10.3390/cryst15040309

Chicago/Turabian Style

Zhang, Zhongzheng, Chunhong Zhang, Xinmao Qin, and Wanjun Yan. 2025. "Research on the Mechanical and Photoelectric Properties Regulation of the New-Type Ceramic Material Ta2AlC" Crystals 15, no. 4: 309. https://doi.org/10.3390/cryst15040309

APA Style

Zhang, Z., Zhang, C., Qin, X., & Yan, W. (2025). Research on the Mechanical and Photoelectric Properties Regulation of the New-Type Ceramic Material Ta2AlC. Crystals, 15(4), 309. https://doi.org/10.3390/cryst15040309

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