Next Article in Journal
Influence of Applied Load and Sliding Velocity on Tribocorrosion Behavior of 7075-T6 Aluminum Alloy
Next Article in Special Issue
Mechanical Properties and Microstructural Evolution of Ti-25Nb-6Zr Alloy Fabricated by Spark Plasma Sintering at Different Temperatures
Previous Article in Journal
Automatic Control of Hot Metal Temperature
Previous Article in Special Issue
Achieving Tunable Microwave Absorbing Properties by Phase Control of NiCoMnSn Alloy Flakes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effects of Vacancies on the Structural, Elastic, Electronic and Thermodynamic Properties of C11b-VSi2 by First-Principles Calculations

Faculty of Material Science and Engineering, Kunming University of Science and Technology, Kunming 650093, China
*
Author to whom correspondence should be addressed.
Metals 2022, 12(10), 1625; https://doi.org/10.3390/met12101625
Submission received: 25 August 2022 / Revised: 13 September 2022 / Accepted: 23 September 2022 / Published: 28 September 2022

Abstract

:
The effects of V and Si vacancies on structural stability, elastic properties, brittleness-toughness transition, Debye temperature and electronic properties of tetragonal C11b-VSi2 are investigated using the first-principles calculations. The vacancy formation energy and phonon dispersions confirm that perfect C11b-VSi2 and C11b-VSi2 with different atomic vacancies are thermodynamically and dynamically stable. The C11b-VSi2 with V-atom vacancies is more stable than that with Si-atom vacancies. The introduction of different atomic vacancies enhances the elastic modulus and its anisotropy of C11b-VSi2. The electron density difference and densities of state of perfect VSi2 and VSi2 with different vacancies are calculated, and the chemical bonding properties of perfect VSi2 and VSi2 with vacancies are discussed and analyzed. Additionally, the results show that the chemical bond strength of VSi2 is enhanced by the introduction of vacancies. Finally, Debye temperatures of perfect VSi2 and VSi2 with vacancies are also calculated.

1. Introduction

Transition metal (TM) disilicides have excellent properties such as high strength, high melting point, high temperature creep resistance and good oxidation resistance in high-temperature environments, and have a wide range of promising applications in high temperature devices and semiconductors [1,2,3,4]. Moreover, transition metal disilicide is a promising high-temperature structural material because of its excellent overall performance at high temperatures. In addition, with the rapid development of the aerospace industry, the demand for structural materials with low density and good high-temperature properties is becoming stronger. Although the strength of transition metal disilicide can meet the requirements of high-temperature materials, its inherent brittleness seriously affects its application. Currently, Xiang et al. [5] investigated the high thermal emission of transition metal disilicides by first-principles calculations. The results showed that the addition of SiO2 significantly improved their high-temperature properties. Zhang et al. [6] systematically investigated the phase stability, mechanical properties, and electronic structure of transition metal disilicides by first-principles. The results show that for tetragonal TiSi2, NiSi2, CuSi2 and ZnSi2, they are either thermodynamically unstable or mechanically unstable.
In addition, TMSi2 (TM = V, Nb and Ta) disilicide is an important member of the transition metal disilicide family. Additionally, many of its properties have been extensively studied, such as magnetization [7], electronic structure and mechanical properties. In previous studies, we could determine that TMSi2 (TM = V, Nb and Ta) had thermodynamic stability and optical anisotropy [8]. Moreover, VSi2 is of interest because of its more excellent optical anisotropy [9]. In previous studies, the structural stability, mechanical and thermodynamic properties of VSi2 have been investigated by experimental and theoretical methods, and it has been shown that VSi2 has excellent comprehensive properties [10,11,12,13]. It is well known that VSi2 consists of two crystal structures, namely, the tetragonal C11b structure and the hexagonal C40 structure. However, for most of the TMSi2 disilicides, the C40 structure is more stable than the C11b structure, and the formation energy differences in the C11b/C40 competition are less than 0.1 eV/atom [14], indicating that the phase stability of C11b is very close to that of C40. The C40 structure compound has desirable physical properties such as high melting point and low density, which are more attractive in the field of microelectronic devices. The C11b structure compound has good overall mechanical properties at higher temperatures, but it has brittle fractures at room temperature [15,16]. Therefore, it is necessary to investigate the overall properties of C11b-VSi2 in order to reduce its brittle behavior and obtain high strength. Different crystal structures and various defects (including vacancies, dislocations, etc.) combine to determine the elastic properties of the material [17,18]. Therefore, vacancies play a crucial role in the mechanical properties of high-temperature materials. At present, the physical properties of VSi2 containing vacancy defects are less studied. Therefore, it is essential to study the effect of vacancies on the overall properties of high-temperature alloys [19].
In this article, we study the effect of different atomic vacancies on the mechanical performance of the tetragonal C11b-VSi2 by first-principles calculations based on density functional theory (DFT) [20]. This theory has been widely used to predict the electronic, mechanical, and thermal properties of solids [21,22,23,24,25,26,27,28]. On the basis of the VSi2 vacancy model, we have studied five different vacancies. To check the stability of the vacancies, the vacancy formation energy and phonon dispersion were calculated for the different vacancies. The effects of vacancies on the elastic properties, tough and brittle behavior and electronic properties of the material were investigated in detail, and it was concluded that vacancies can improve the brittle behavior of C11b-VSi2.

2. Calculation Method

In this paper, based on density functional theory (DFT), first-principles calculations are performed on the structure, elastic properties, electronic properties and chemical bonds of perfect C11b-VSi2 and C11b-VSi2 with different vacancies through the CASTEP [29] code. The interactions between the ionic nuclei and valence electrons are described in terms of a ultrasoft pseudopotentials. The electronic configurations of the pseudo-atoms are V 3s23p63d34s2 and Si 3s23p2, respectively. To verify that the total energy in the ground state is convergent, a plane wave basis is set to an electronic wave function with a cut-off energy of 500 eV. All lattice properties, atomic locations and internal coordinates in the system are entirely loosened during structural optimization. Using the PHONOPY code [30], we calculated the phonon frequencies for ideal C11b-VSi2 and VSi2 including varying vacancies to assess the dynamic stability.

3. Results and Discussion

3.1. Structural Properties for C11b-VSi2 with Different Atomic Vacancies

Here, VSi2 is a tetragonal (C11b) structure. To investigate the relationship between different vacancies and properties, we constructed a crystal structure containing nine V atoms and ten Si atoms as shown in Figure 1 and discussed C11b-VSi2 with different vacancies, such as V-va1, V-va2, Si-va1 and Si-va2 [31]. By discussing the forming energies of perfect C11b-VSi2 and different atomic vacancies C11b-VSi2, the thermodynamic stability of C11b-VSi2 containing vacancies is studied. The vacancy formation energy (Ef), which is the energy necessary to generate a vacancy and is dependent on the relationship between the atom and its nearby atoms whenever the element is removed, is used to determine the thermodynamic stability of vacancies [32]. Only when the formation energy is negative, it is known that a solid material is thermodynamically stable. In particular, the lower the value of the vacancy formation energy, the better the thermodynamic stability [33]. When atomic vacancies are present, C11b-VSi2 with different atomic vacancies have different levels of thermodynamic stability. Therefore, to investigate the structural stability of C11b-VSi2, we estimated the atomic vacancy formation energy as follows:
E f = E V a c M E P e r f T o t a l + μ M
The total energy of C11b-VSi2 with an M (M = V and Si) vacancy and a C11b-VSi2 perfect crystal without a vacancy are E V a c M and E P e r f T o t a l , respectively. μM is the chemical potential of the M atom that has been eliminated.
The vacancy formation energies and lattice parameters of perfect C11b-VSi2 and C11b-VSi2 containing different atomic vacancies are listed in Table 1, as well as the other theoretical results [10]. Because the vacancy formation energies in C11b-VSi2 are less than zero, these vacancies are thermodynamically stable in the ground state [34]. Moreover, with the introduction of vacancies, the vacancy formation energy of C11b-VSi2 containing vacancies is slightly larger compared with that of the perfect C11b-VSi2. It shows that the introduction of vacancies increases the thermodynamic instability of C11b-VSi2. The calculated vacancy formation energies of V vacancies are smaller than those of Si vacancies, as shown in Table 1, indicating that the V vacancies are more thermodynamically stable than the Si-vacancies. As a result, C11b-VSi2 is more inclined to form V vacancies.
Moreover, the lattice parameters of perfect C11b-VSi2 and C11b-VSi2 containing different vacancies are discussed in depth, which have also been discussed in NbSi2 [35]. In general, the lattice parameters of C11b-VSi2 containing vacancies are slightly smaller than that of perfect C11b-VSi2 along the a-axis and the b-axis, but slightly larger than that of perfect C11b-VSi2 along the c-axis. The differences in the lattice parameters determine the elastic properties. For V vacancies, the lattice constants of V-va1 are approximately equal to those of V-va2. Additionally, C11b-VSi2 with V vacancies have larger values of lattice parameters a and b than Si vacancies; however, the c values are smaller than those of Si vacancies. For Si vacancies, the lattice constant values for Si-va2 vacancies are larger than those for Si-va1 in the a and b directions, and the c values are approximately equal. The position and atomic type of the removed atoms lead to differences in the variation of the lattice constants.
The computed phonon dispersion of perfect C11b-VSi2 and C11b-VSi2 with vacancies in the Brillouin zone along the high symmetric direction is shown in Figure 2. The phonon spectrum is an important indicator to judge whether the system is stable [36]. If the phonon frequencies are all above the zero point, it means that the material does not have imaginary frequencies, indicating that the material is stable [37]. From Figure 2, the phonon frequencies of both perfect C11b-VSi2 and C11b-VSi2 in the presence of vacancies have no imaginary frequencies, indicating that they are dynamically stable. This yields results that are consistent with those obtained for the vacancy formation energy.

3.2. Elastic Properties

Before analyzing the elastic properties of C11b-VSi2 with various vacancies, it is necessary to consider the mechanical stability of the material [38]. The elastic constants (Cij) and the elastic flexibility matrix (Sij) are calculated for perfect C11b-VSi2 and C11b-VSi2 containing different vacancies in Table 2 and Table 3, respectively. The generalized mechanical stability criterion for the tetragonal system is C11 > 0, C33 > 0, C44 > 0, C66 > 0, C11 > C12, C11 + C33 − 2C13 > 0, 2C11 + C33 + 2C12 + 4C13 > 0 [39]. Clearly, the perfect C11b-VSi2 and C11b-VSi2 containing different vacancies satisfy the stability criterion for all elastic constants. Therefore, they have mechanical stability.
The linear compression resistance along the a-axis, b-axis, and c-axis is represented by C11, C22, and C33, respectively [40]. Table 2 shows that the C11 values for perfect C11b-VSi2 and C11b-VSi2 containing atomic vacancies are generally lower than the C33 values, indicating that the deformation resistance along the a-axis is lower than the deformation resistance along the c-axis. The C11 values of C11b-VSi2 with different vacancies are larger than those of perfect C11b-VSi2, yet the C33 values are lower than those of perfect C11b-VSi2. It shows that the vacancies created by the removal of atoms increase the resistance to deformation along the a-axis but weaken resistance to deformation along the c-axis. Additionally, since the C11 value of C11b-VSi2 containing Si vacancies is larger than that of C11b-VSi2 containing V vacancies, the deformation resistance produced by the removal of Si atoms is larger than that produced by the removal of V atoms in the direction along the a-axis. In addition to this, it is well known that C44 and C66 are correlated to shear modulus, with larger values of C44 and C66 corresponding to larger shear modulus. Table 2 shows that the C44 and C66 values of C11b-VSi2 with different vacancies are larger than those of perfect C11b-VSi2, indicating that the atom vacancies significantly enhance the shear deformation resistance of C11b-VSi2. Moreover, the V-va2 vacancies has the largest C33, C44 and C66 values compared to other vacancy types, which implies that C11b-VSi2 with V-va2 vacancies have a greater resistance to deformation.
The mechanical properties of C11b-VSi2 containing these vacancies are investigated in order to establish the link between vacancies and mechanical performance. The elastic moduli (including bulk modulus B, Poisson’s ratio v, Young’s modulus E, and shear modulus G) of perfect C11b-VSi2 and C11b-VSi2 with different atomic vacancies for the C11b structure obtained by the Voigt–Reuss–Hill method are presented in Table 4, where BV (GV) and BR (GR) are B (G) in the Voigt and Reuss approximations, respectively, and the expressions are as follows [41,42,43,44,45]:
B = B V + B R 2
G = G V + G R 2
E = 9 GB G + 3 B
ν = 3 B E 6 B
H V = 2 ( G 3 B 2 ) 0.585 3
The ratio of bulk modulus to shear modulus (B/G) validates the solid material’s brittleness/ductility [46]. If the B/G ratio is higher than 1.75, the solid material will exhibit ductility. Otherwise, the material will exhibit brittleness. The higher the B/G value, the more ductile the solid material is [47]. Based on the B/G ratios of C11b-VSi2 with varying vacancies as well as perfect C11b-VSi2 shown in Table 4, the B/G of perfect C11b-VSi2 is 1.642, which is less than 1.75, showing that C11b-VSi2 is brittle. Meanwhile, with the introduction of vacancies, the calculated B/G values of C11b-VSi2 containing these vacancies generally decrease, except the B/G value of V-va1 (1.668). Notably, the B/G values of C11b-VSi2 with V vacancies are all larger than those of C11b-VSi2 with Si vacancies, and thus, C11b-VSi2 containing V vacancies have less brittleness.
Bulk modulus is a macroscopic property of a material that reflects its resistance to external compression, i.e., incompressibility, as is widely known. As seen in Table 4, the bulk modulus of C11b-VSi2 containing different vacancies is larger than that of perfect C11b-VSi2, indicating that the removal of atoms instead makes the material more incompressible. The force that resists form change under shear stress is known as the shear modulus (G). Shear stress is more closely related to hardness than bulk modulus, and shear modulus is a more appropriate predictor of hardness [48]. In Table 4, C11b-VSi2 with the introduction of V-va1 vacancy has the smallest shear modulus and is smaller than perfect C11b-VSi2. Therefore, its ductility is slightly better than that of perfect C11b-VSi2. Additionally, all the remaining types of vacancies increase the shear modulus, compared to perfect C11b-VSi2. Thus, the hardness values of C11b-VSi2 with V-va2, Si-va1, and Si-va2 increase, and this variation of hardness with vacancy has been found in chromium silicide [49]. In addition, the stiffness of a solid can be described using Young’s modulus. The stiffness of a solid is proportional to its Young’s modulus. As can be seen from Table 4, the introduction of atomic vacancies increases the Young’s modulus of C11b-VSi2, with the largest being for Si-va1 vacancies and only those containing V-va1 vacancies having lower Young’s modulus. Accordingly, removing atoms increases the elastic stiffness of VSi2. It is congruent with the results of the elastic constants. From the point of view of interatomic interactions, elastic modulus is closely related to the interaction between atoms. The increased elastic modulus of C11b-VSi2 containing different vacancies, except for the V-va1 vacancy, can be tentatively judged that the removal of atoms enhances the atomic interaction of C11b-VSi2.
Poisson’s ratio is a well-known method for determining a solid’s stability under shear deformation [50]. Solids having a Poisson’s ratio of −1 to 0.5 are relatively stable under shear deformation in general. Furthermore, a larger Poisson’s ratio indicates that the solid is more malleable. As shown in Table 4, the Poisson’s ratios of perfect C11b-VSi2 and C11b-VSi2 containing different vacancies is in the range of 0.229~0.25, which is between −1 and 0.5, indicating that they are stable solids. Among them, VSi2 containing V-va1 vacancies has the largest Poisson’s ratio, indicating that it has better plasticity than the other compounds. It also corroborates that the Vickers hardness of VSi2 containing V-va1 vacancies is the smallest among them. Poisson’s ratio also can be used to estimate the solid’s brittleness and ductility. If the solid has a ν > 0.33, the material is ductile; otherwise, it is fragile. From the calculations in Table 4, ν < 0.33 for perfect C11b-VSi2 and C11b-VSi2 with different vacancies; thus, they are brittle. This result agrees well with that of B/G.

3.3. Elastic Anisotropy

The elastic anisotropy is an important index reflecting the mechanical anisotropy of materials and plays a very large role in the generation of microcracks. To better describe the elastic anisotropy, in this paper, we use elastic anisotropy indices such as the universal anisotropy index (AU), compression and shear anisotropy percentages (Acomp and Ashear), and shear anisotropy factors (A1, A2, and A3) to investigate the elastic anisotropy [51]. The calculation equations are as follows:
A U = B V B R + 5 G V G R 6
A comp = B V B R B V + B R × 100 %
A shear = G V G R G V + G R × 100 %
Furthermore, due to the fact that the VSi2 we study in this paper is a tetragonal crystal structure, considering its shear anisotropy factor:
A 1 = 4 C 44 C 11 + C 33 2 C 13
A 2 = 4 C 55 C 22 + C 33 2 C 23
A 3 = 4 C 66 C 11 + C 22 2 C 12
A1, A2, and A3 represent the degree of shear anisotropy corresponding to the (100), (010) and (001) planes, respectively. For these elastic anisotropy indices, if A1 = A2 = A3 = 1 and AU = Acomp = Ashear = 0 [52], the solid shows elastic isotropy; otherwise, it is anisotropic. At the same time, when the solid has larger values of AU, Acomp and Ashear, it has higher elastic anisotropy.
The calculated elastic anisotropic indices are listed in Table 5. From Table 5, both perfect C11b-VSi2 and C11b-VSi2 with different vacancies are anisotropic, since their elastic anisotropy indices deviate from 0. The greater the deviation, the greater the anisotropy [53]. It is obvious that the changes corresponding to Acomp and Ashear are different. Ashear increases with the introduction of vacancies. However, Acomp is decreasing. This may be due to the difference in the variation of bulk and shear moduli. Therefore, using Acomp and Ashear alone to evaluate the material elastic anisotropy has limitations. On the contrary, AU is more accurate to evaluate the elastic anisotropy by considering both bulk modulus and shear modulus. As can be seen from Table 5, the AU values of C11b-VSi2 containing different vacancies are larger than those of VSi2, and in contrast, the AU values of C11b-VSi22 containing V vacancies increase sharply. It indicates that the introduction of vacancies improves the anisotropy of the material, especially the V vacancies. Among them, the elastic anisotropy of vacancy V-va1 is the largest. In addition, the shear anisotropy indices A1, A2 and A3 are not equal to 1, indicating that they are all anisotropic. The introduction of the vacancy increases A1, A2 and A3, meaning that the vacancy increases the shear anisotropy of C11b-VSi2 in the (100), (010) and (001) plane all. For A3, the increase is the largest. It shows that the vacancy enhances the shear anisotropy of C11b-VSi2 in the (001) plane. the increase in V-va1 is particularly significant.
The elastic anisotropy of the crystal can also be visualized by the three-dimension (3D) surface construction diagram of the elastic modulus. When the crystal is isotropic, the 3D surface construction diagram is perfectly spherical. On the contrary, it is anisotropic. In this work, we focus on the elastic anisotropy of the bulk modulus B and Young’s modulus E through the 3D surface construction diagram. For the tetragonal structure, the inverse of the bulk modulus B and Young’s modulus E are calculated as follows [54]:
1 B = ( S 11 + S 12 + S 13 ) - ( S 11 + S 12 - S 13 - S 33 ) l 3 2
1 E = S 11 ( l 1 4 + l 2 4 ) + ( 2 S 13 + S 44 ) ( l 1 2 l 3 2 + l 2 2 l 3 2 ) + S 33 l 3 4 + ( 2 S 12 + S 66 ) l 1 2 l 2 2
Here, Sij is the elastic compliance constant listed in Table 3, and l1, l2, and l3 are the direction cosines.
Figure 3 and Figure 4 show the 3D surface construction diagrams of the bulk modulus and Young’s modulus, respectively. When the deviation of the 3D surface construction diagram from the spherical shape is greater, the anisotropy of the solid is greater [55]. As can be seen from Figure 3, there is no significant difference between the 3D diagrams of bulk modulus of C11b-VSi2 containing vacancies, then it is necessary to consider the 3D diagram of Young’s modulus. It is obvious from Figure 4 that the 3D diagrams of Young’s modulus of C11b-VSi2 containing different vacancies and perfect C11b-VSi2 are non-spherical. Compared with the 3D diagrams of bulk modulus, the 3D diagrams of Young’s modulus are more irregular in shape, indicating that they are all anisotropic, and the anisotropic feature is greater than that of the bulk modulus. Moreover, with the introduction of vacancies, the deviation of 3D diagrams of Young’s modulus from the spherical shape is greater, indicating that the different atomic vacancies enhance the anisotropy of Young’s modulus. This result is in good agreement with the results corresponding to AU in Table 5.
However, the 3D construction diagrams in Figure 4 do not clearly show the subtle differences in the elastic modulus anisotropy of C11b-VSi2 containing different vacancies. Therefore, to see more details of the elastic anisotropy, the two-dimension (2D) projections of bulk modulus and Young’s modulus in the (001) and (100) planes, which have been employed successfully [56], for both perfect C11b-VSi2 and C11b-VSi2 containing different atomic vacancies are shown in Figure 5. From Figure 5, in general, the introduction of vacancies increases the anisotropy of the elastic modulus. For the Young’s modulus, Figure 5a shows that the shape of C11b-VSi2 containing different atomic vacancies in the (001) plane is more deviated from circular than that of perfect C11b-VSi2. It means that the vacancies make Young’s modulus more anisotropic. As shown in Figure 5c, the 2D projection of C11b-VSi2 containing Si atomic vacancies in the (100) plane is approximately or slightly anisotropic to that of perfect C11b-VSi2. In contrast, C11b-VSi2 containing V-atom vacancies is more obviously irregular in shape and more anisotropic. This result is in agreement with the one corresponding to AU. For the bulk modulus, in the (001) plane, the shapes of the graphs are all circular and the crystals show no significant anisotropy. However, in the (100) plane, the shape of C11b-VSi2 containing different vacancies deviates significantly from a circle and shows a higher anisotropy.
Table 6 shows the elastic moduli in the [100], [010] and [001] directions obtained according to Figure 5. It is obvious from Table 6 that the values of Young’s modulus and bulk modulus in the [100] and [010] directions are smaller than those in the [001] direction because the value of C11 is smaller than that of C33, making the crystal easier to compress along the a-axis [57].

3.4. Electronic Properties

As mentioned above, the reason for the change in the elastic modulus of C11b-VSi2 containing atomic vacancies is related to the inter-charge interactions, where the type and position of the atoms determine the inter-charge interactions [58]. The removal of V and Si atoms in C11b-VSi2 due to inter-charge interactions changes the electron equilibrium concentration between adjacent atoms and alters the chemical bonding in VSi2. There are two different types of chemical bonds in perfect C11b-VSi2, the V-Si bond and the Si-Si covalent bond. The bond lengths of its Si-Si covalent and V-Si bonds were calculated to be 2.529 Å and 2.579 Å, respectively, which is consistent with earlier theoretical predictions. The bond lengths of Si-Si covalent and V-Si bonds for perfect C11b-VSi2 and C11b-VSi2 with different vacancies are shown in Table 7. As shown in Table 7, the insertion of Si vacancies boosts the charge interaction between the V atom and the Si atom while weakening the charge interaction between the Si atom and the other Si atom. As a result, the single cell cohesion energy is weakened due to electron collapse, leading to lattice contraction. This is the major reason why Si vacancies have a greater elastic modulus than perfect C11b-VSi2.
For the V-atom vacancy, the presence of the V atom vacancy shortens the bond length of both the V-Si bond and the Si-Si covalent bond. In other words, the removal of the V atom enhances the interatomic charge interactions. We suggest that the loss of the V-Si bond on the shear surface causes the enhancement of the elastic modulus.
The electron density difference can be sued to further investigate electron transfer in chemical bonding [59]. Figure 6 shows the electron density difference for perfect C11b-VSi2 and C11b-VSi2 with V and Si atomic vacancies. In each graph, the electron density difference ranges from −0.1e/Å3 to 0.1e/Å3. The blue color implies the maximum localization of electrons in this region, and the red color indicates the maximum delocalization of electrons in this region. In Figure 6a, intact V-Si bonds and Si-Si covalent bonds in perfect C11b-VSi2 are observed. The removal of atoms clearly disrupts the localized hybridization between adjacent atoms compared to the figure with atomic vacancies. By comparison, we find that the electron delocalization of V vacancies and Si vacancies is weaker than that of perfect C11b-VSi2. In other words, the localized hybridization between V and Si atoms is improved by the removed atoms, which has been confirmed in TaSi2 that the removals of Ta and Si can enhance the electron hybridization between Ta and Si [60]. As can be seen in Figure 6, for C11b-VSi2 containing vacancies, the inter-atomic charge interactions are stronger than the corresponding charge interactions for perfect C11b-VSi2. This is the reason that the elastic modulus of C11b-VSi2 containing vacancies increases.
In Figure 6b,c, for Si vacancies, the removal of Si atoms enhances the local hybridization between V and Si atoms, but the Si-Si atom interactions are weakened. This is consistent with the results obtained from the discussion of bond lengths above. Thus, the enhanced elastic modulus comes from the electronic bonding properties. From Figure 6d,e, the removal of V atoms when V vacancies are present also results in enhanced Si-Si inter-atomic interactions.
We calculated the total and partial densities of states (DOS and PDOS) to discuss the bonding properties of perfect C11b-VSi2 and C11b-VSi2 containing different atomic vacancies in Figure 7. DOS denotes the number of electron states per unit energy interval when the electron energy levels are quasi-continuously distributed [61,62]. The dashed line with zero energy represents the Fermi energy level [63]. The valley at the Fermi energy level is called the pseudogap, which indicates the presence of a stable phase. From the comparison in Figure 7, the pseudogap of C11b-VSi2 with V vacancies is the smallest, proving that V vacancies are more stable than Si vacancies. This result is in line with the vacancy formation energy conclusion. The TDOS profile of C11b-VSi2 around the Fermi energy level is predominantly from the V-3d state and the Si-3p state, showing substantial hybridization between V and Si atoms, as seen in Figure 7. Figure 7a shows that the Si-3p state is divided into two parts by the Si-3s state, indicating the formation of hybridization between Si and Si atoms in VSi2. However, the PDOS profile resulting from the removal of V and Si atoms is slightly different from that of the perfect C11b-VSi2. The energy of the V vacancies is determined by a tiny change around the Fermi energy level. As can be seen in Figure 7b,c, for the V vacancy, the PDOS produces some small peaks near the Fermi energy level. The charge energy of V and Si atoms increases compared to that of perfect C11b-VSi2. The introduction of Si vacancies has a similar feature. The DOS profile of C11b-VSi2 with Si vacancies is similar to that of C11b-VSi2 with V vacancies, while more electrons are transferred from the lower energy region to the Fermi energy level compared with the density of states of V vacancies. It leads to stronger local hybridization between the Si-3s and Si-3p states, forming Si-Si covalent bonds. The loss of Si atoms, in particular, causes a charge transfer from the Si-3s state to the Si-3p state.

3.5. Debye Temperature

The Debye temperature (θD) is a fundamental thermodynamic parameter that is affected by melting point, coefficient of thermal expansion, specific heat, and other properties, and describes the lattice vibrations and changes in specific heat [64]. In order to assess the overall performance of high-temperature alloys, the Debye temperatures of perfect C11b-VSi2 and C11b-VSi2 with different atomic vacancies need to be considered [65]. Under the Debye model, the Debye temperature of a solid is given by the modulus of elasticity, which is calculated from the sound velocity, and the expression for the Debye temperature is as follows:
θ D = h K B [ 3 n 4 π ( N A ρ M ) ] 1 3 v m
In the formula, kB is Boltzmann’s constant, n is the total number of atoms in the molecule, NA is Avogadro’s constant, h is Planck’s constant, ρ is the density, and M is the molecular weight.
The longitudinal and transverse sound velocities are denoted by νl and νt, respectively, and the formulas are as follows:
v l = [ ( B + 4 G 3 ) / ρ ] 1 2
v t = ( G ρ ) 1 2
The mean sound velocity can be calculated from νl and νt with the following equations:
v m = [ 1 3 ( 2 v t 3 + 1 v l 3 ) ] - 1 3
Table 8 lists the density, longitudinal sound velocity, transverse sound velocity, mean sound velocity and Debye temperature for perfect C11b-VSi2 and C11b-VSi2 with different vacancies. It is known that the properties of the chemical bonds determine the Debye temperature. The higher the Debye temperature, the higher the interatomic forces and the higher the chemical bond strength. As can be seen from Table 8, the Debye temperature containing V vacancies and Si vacancies is larger than that of perfect C11b-VSi2, indicating that the introduction of vacancies enhances the chemical bond strength. Among them, the highest Debye temperature for the Si-va1 vacancy results in the strongest chemical bond strength. This is consistent with the conclusions obtained in the previous section.

4. Conclusions

In summary, we used first-principle calculations to study the structural properties, thermal stability, elastic properties, electronic properties and Debye temperature of perfect C11b-VSi2 and C11b-VSi2 with different vacancies and have concluded the following:
(1) The introduction of V vacancies and Si vacancies in C11b-VSi2 is thermodynamically stable at the ground state. Moreover, the V vacancies are more thermodynamically stable than the Si vacancies in C11b-VSi2.
(2) The introduction of vacancies enhanced the bulk modulus, shear modulus and Young’s modulus of C11b-VSi2, which significantly improved the mechanical behavior of C11b-VSi2.
(3) The elastic anisotropy results indicate that the introduction of vacancies enhances the elastic anisotropy of C11b-VSi2, and the V-atom vacancies are the most pronounced.
(4) The difference in electron density difference and density of states between perfect C11b-VSi2 and C11b-VSi2 with different vacancies suggests that the introduction of vacancies enhances the interactions between charges.
(5) Calculations on the Debye temperature show that the introduction of vacancies enhances the chemical bond strength of C11b-VSi2.

Author Contributions

Conceptualization, writing—original draft preparation, methodology, S.X.; visualization, validation, M.P.; data curation, formal analysis, L.S.; supervision, project administration, funding acquisition, Y.D. All authors have read and agreed to the published version of the manuscript.

Funding

This study is supported by the Yunnan Ten Thousand Talents Plan Young & Elite Talents Project under Grant no. YNWR-QNBJ-2018-044.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Petrovic, J.J.; Vasudevan, A.K. Key developments in high temperature structural silicides. Mater. Sci. Eng. A 1999, 261, 1–5. [Google Scholar] [CrossRef]
  2. Gottlieb, U.; Lasjaunias, J.C.; Laborde, O.; Thomas, O.; Madar, R. Low temperature specific heat measurements of VSi2, NbSi2 and TaSi2. Appl. Surf. Sci. 1993, 73, 232–236. [Google Scholar] [CrossRef]
  3. Chen, Y.; Kolmogorov, A.N.; Pettifor, D.G.; Shang, J.X.; Zhang, Y. Theoretical analysis of structural stability of TM5Si3 transition metal silicides. Phys. Rev. B 2010, 82, 184104. [Google Scholar] [CrossRef]
  4. Vethaak, T.D.; Gustavo, F.; Farjot, T.; Kubart, T.; Gergaud, P.; Zhang, S.-L.; Lefloch, F.; Nemouchi, F. Superconducting V3Si for quantum circuit applications. Microelectron. Eng. 2021, 111570, 244–246. [Google Scholar] [CrossRef]
  5. Xiang, H.; Dai, F.; Zhou, Y. Secrets of high thermal emission of transition metal disilicides TMSi2 (TM = Ta, Mo). J. Mater Sci. Technol. 2021, 89, 114–121. [Google Scholar] [CrossRef]
  6. Zhang, X.; Wang, Z.; Qiao, Y. Prediction of stabilities, mechanical properties and electronic structures of tetragonal 3d transition metal disilicides: A first-principles investigation. Acta Mater. 2011, 59, 5584–5592. [Google Scholar] [CrossRef]
  7. Gottlieb, U.; Sulpice, A.; Madar, R.; Laborde, O. Magnetic susceptibilities of VSi2, NbSi2 and TaSi2 single crystals. J. Phys. Condens. Matter. 1993, 5, 8755. [Google Scholar] [CrossRef]
  8. Laborde, O.; Bossy, J.; Affronte, M.; Schober, H.; Gottlieb, U. Some properties of the phonon spectra of transition metal disilicides VSi2, NbSi2, and TaSi2. Solid State Commun. 2003, 126, 415–419. [Google Scholar] [CrossRef]
  9. Yang, C.; Wu, Y.; Duan, Y. Theoretical predictions of the electronic, optical and thermodynamic properties of the C40-type TMSi2 (TM = V, Nb and Ta) disilicides. Mater. Today Commun. 2022, 30, 103115. [Google Scholar] [CrossRef]
  10. Káňa, T.; Šob, M.; Vitek, V. Ab initio study of phase transformations in transition-metal disilicides. Intermetallics 2011, 19, 919–926. [Google Scholar] [CrossRef]
  11. Yuge, K.; Kishida, K.; Inui, H.; Koizumi, Y.; Hagihara, K.; Nakano, T. Cr segregation at C11b/C40 interface in MoSi2-based alloys: A first-principles study. Intermetallics 2013, 42, 165–169. [Google Scholar] [CrossRef]
  12. Kishida, K.; Nakatsuka, S.; Nose, H.; Inui, H. Room-temperature deformation of single crystals of transition-metal disilicides (TMSi2) with the C11b (TM = Mo) and C40 (TM = V, Cr, Nb and Ta) structures investigated by micropillar compression. Acta Mater. 2022, 223, 117468. [Google Scholar] [CrossRef]
  13. Ghadami, S.; Taheri-Nassaj, E.; Baharvandi, H.R.; Ghadami, F. Effect of in situ VSi2 and SiC phases on the sintering behavior and the mechanical properties of HfB2-based composites. Sci. Rep. 2020, 10, 16540. [Google Scholar] [CrossRef] [PubMed]
  14. Carlsson, A.E.; Meschter, P.J. Energetics of C11b, C40, C54, and C49 structures in transition-metal disilicides. J. Mater. Res. 1991, 6, 1512–1517. [Google Scholar] [CrossRef]
  15. Chu, F.; Lei, M.; Maloy, S.A.; Petrovic, J.J.; Mitchell, T.E. Elastic properties of C40 transition metal disilicides. Acta Mater. 1996, 44, 3035–3048. [Google Scholar] [CrossRef]
  16. Wan, B.; Xiao, F.; Zhang, Y.; Zhao, Y.; Wu, L.; Zhang, J.; Gou, H.Y. Theoretical study of structural characteristics, mechanical properties and electronic structure of metal (TM = V, Nb and Ta) silicides. J. Alloy Compd. 2016, 681, 412–420. [Google Scholar] [CrossRef]
  17. Nunes, C.A.; de Lima, B.B.; Coelho, G.C.; Suzuki, P.A. Isothermal section of the V-Si-B system at 1600 °C in the V-VSi2-VB region. J. Phase Equilib. Diff. 2009, 30, 345–350. [Google Scholar] [CrossRef]
  18. Inui, H.; Moriwaki, M.; Yamaguchi, M. Plastic deformation of single crystals of VSi2 and TaSi2 with the C40 structure. Intermetallics 1998, 6, 723–728. [Google Scholar] [CrossRef]
  19. Zhao, X.H.; Wei, X.N.; Tang, T.Y.; Xie, Q.; Gao, L.K.; Lu, L.M.; Hu, D.Y.; Li, L.; Tang, Y.L. Theoretical prediction of the structural, electronic and optical properties of vacancy-ordered double perovskites Tl2TiX6 (X = Cl, Br, I). J. Solid State Chem. 2022, 305, 122684. [Google Scholar] [CrossRef]
  20. Hu, B.; Yao, B.; Wang, J.; Liu, Y.; Wang, C.; Du, Y.; Yin, H. The phase equilibria of the Ti-V-M (M = Si, Nb, Ta) ternary systems. Intermetallics 2020, 118, 106701. [Google Scholar] [CrossRef]
  21. Wang, Y.; Wu, Y.; Lu, Y.; Wang, X.; Duan, Y.; Peng, M. Theoretical insights to elastic and thermal properties of WB4 tetraborides: A first-principles calculation. Vacuum 2022, 196, 110731. [Google Scholar] [CrossRef]
  22. Yang, A.; Duan, Y.; Peng, M. Effects of temperature and pressure on the mechanical and thermodynamic properties of high-boride WB4 from first-principles predictions. Mater. Today Commun. 2022, 30, 103187. [Google Scholar] [CrossRef]
  23. Sun, Y.; Yang, A.; Duan, Y.; Shen, L.; Peng, M.; Qi, H. Electronic, elastic, and thermal properties, fracture toughness, and damage tolerance of TM5Si3B (TM = V and Nb) MAB phases. Int. J. Refract. Met. Hard Mater. 2022, 103, 105781. [Google Scholar] [CrossRef]
  24. Yang, A.; Duan, Y.; Yi, J.; Li, C. Theoretical insights into anisotropies in elastic and thermal properties of ternary β-M4AlN3 (M = V, Nb, Ta) nitrides by first-principles calculations. Chem. Phys. Lett. 2021, 783, 139088. [Google Scholar] [CrossRef]
  25. Lu, Y.; Yang, A.; Duan, Y.; Peng, M. Structural stability, electronic and optical properties of MAX-phase ternary nitrides β-TM4AlN3 (TM = V, Nb, and Ta) using the first-principles explorations. Vacuum 2021, 193, 110529. [Google Scholar] [CrossRef]
  26. Lu, Y.; Duan, Y.; Peng, M.; Yi, J.; Li, C. First-principles calculations of electronic, optical, phononic and thermodynamic properties of C40-type TMSi2 (TM = Cr, Mo, W) disilicides. Vacuum 2021, 191, 110324. [Google Scholar] [CrossRef]
  27. Wu, Y.; Wang, X.; Wang, Y.; Duan, Y.; Peng, M. Insights into electronic and optical properties of AGdS2 (A = Li, Na, K, Rb and Cs) ternary gadolinium sulfides. Opt. Mater. 2021, 114, 110963. [Google Scholar] [CrossRef]
  28. Qu, D.; Li, C.; Bao, L.; Kong, Z.; Duan, Y. Structural, electronic, and elastic properties of orthorhombic, hexagonal, and cubic Cu3Sn intermetallic compounds in Sn-Cu lead-free solder. J. Phys. Chem. Solids 2020, 138, 109253. [Google Scholar] [CrossRef]
  29. Clark, S.J.; Segall, M.D.; Pickard, C.J.; Hasnip, P.J.; Probert, M.I.J.; Refson, K.; Payne, M.C. First principles methods using CASTEP. Z. Krist.-Cryst. Mater. 2005, 220, 567–570. [Google Scholar] [CrossRef]
  30. Togo, A.; Tanaka, I. First principles phonon calculations in materials science. Scr. Mater. 2015, 108, 1–5. [Google Scholar] [CrossRef] [Green Version]
  31. Usseinov, A.; Koishybayeva, Z.; Platonenko, A.; Pankratov, V.; Suchikova, Y.; Akilbekov, A.; Zdorovets, M.; Purans, J.; Popov, A.I. Vacancy defects in Ga2O3: First-principles calculations of electronic structure. Materials 2021, 14, 7384. [Google Scholar] [CrossRef] [PubMed]
  32. Chen, L.J.; Hou, Z.F.; Zhu, Z.Z.; Yang, Y. First-principles calculation of the vacancy formation energies in LiAl. Acta Phys. Sin-Chin. Ed. 2003, 52, 2229–2234. [Google Scholar] [CrossRef]
  33. Guo, G.; Luo, S.; Lai, C.; Ming, B.; You, M.; Liu, X.L.; Huang, Z.Y.; Wei, X.L.; Wang, R.Z.; Zhong, J.X. Substitutional doping effect of C3N anode material: A first principles calculations study. Appl. Surf. Sci. 2022, 571, 151330. [Google Scholar] [CrossRef]
  34. Pan, Y.; Zhang, J.; Jin, C.; Chen, X. Influence of vacancy on structural and elastic properties of NbSi2 from first-principles calculations. Mater. Des. 2016, 108, 13–18. [Google Scholar] [CrossRef]
  35. Bao, W.; Liu, D.; Duan, Y. A first-principles prediction of anisotropic elasticity and thermal properties of potential superhard WB3. Ceram. Int. 2018, 44, 14053–14062. [Google Scholar] [CrossRef]
  36. Bao, L.; Qu, D.; Kong, Z.; Duan, Y. Predictions of structural, electronic, mechanical, and thermodynamic properties of TMBCs (TM = Ti, Zr, and Hf) ceramics. J. Am. Ceram. Soc. 2020, 103, 5232–5247. [Google Scholar] [CrossRef]
  37. Wei, S.; Chou, M.Y. Phonon dispersions of silicon and germanium from first-principles calculations. Phys. Rev. B 1994, 50, 2221–2226. [Google Scholar] [CrossRef]
  38. Qiao, Y.J.; Zhang, H.X.; Hong, C.Q.; Zhang, X.H. Phase stability, electronic structure and mechanical properties of molybdenum disilicide: A first-principles investigation. J. Phys. D Appl. Phys. 2009, 42, 105413. [Google Scholar] [CrossRef]
  39. Bao, W.; Liu, D.; Li, P.; Duan, Y. Structural properties, elastic anisotropies and thermal conductivities of tetragonal LnB2C2 (Ln = Rare Earth) compounds from first-principles calculations. Ceram. Int. 2019, 45, 1857–1867. [Google Scholar] [CrossRef]
  40. Pan, Y.; Pu, D.L.; Yu, E.D. Structural, electronic, mechanical and thermodynamic properties of Cr–Si binary silicides from first-principles investigations. Vacuum 2021, 185, 110024. [Google Scholar] [CrossRef]
  41. Li, B.; Duan, Y.; Peng, M.; Qi, H.; Shen, L.; Wang, X. Theoretical insights on elastic anisotropy and thermal anisotropy of TM5Al3C (TM = Zr, Hf, and Ta) carbides. Vacuum 2022, 200, 110989. [Google Scholar] [CrossRef]
  42. Yang, A.; Xiao, K.; Duan, Y.; Li, C.; Peng, M.; Shen, L. Revealing effects of common nonmetallic impurities on the stability and strength of Cu-Sn solder joints: A first-principles investigation. Vacuum 2022, 200, 110997. [Google Scholar] [CrossRef]
  43. Yang, A.; Duan, Y.; Peng, M.; Shen, L.; Qi, H. Elastic properties, fracture toughness, ideal tensile strength and thermal conductivities of the stable hexagonal WB2, W2B5, WB3 and WB4. Appl. Phys. A 2022, 128, 152. [Google Scholar] [CrossRef]
  44. Yang, A.; Duan, Y.; Li, C.; Yi, J.; Peng, M. Theoretical explorations of structure, mechanical properties, fracture toughness, electronic properties, and thermal conductivity of Ag-doped η′-Cu6Sn5. Intermetallics 2022, 141, 107437. [Google Scholar] [CrossRef]
  45. Chen, X.Q.; Niu, H.; Li, D.; Li, Y. Modeling hardness of polycrystalline materials and bulk metallic glasses. Intermetallics 2011, 19, 1275–1281. [Google Scholar] [CrossRef]
  46. Pu, D.L.; Pan, Y. Influence of high pressure on the structure, hardness and brittle-to-ductile transition of NbSi2 ceramics. Ceram. Int. 2021, 47, 2311–2318. [Google Scholar] [CrossRef]
  47. Pan, Y. First-principles investigation of the new phases and electrochemical properties of MoSi2 as the electrode materials of lithium ion battery. J. Alloys Compd. 2019, 779, 813–820. [Google Scholar] [CrossRef]
  48. Liu, D.; Bao, W.; Duan, Y. Predictions of phase stabilities, electronic structures and optical properties of potential superhard WB3. Ceram. Int. 2019, 45, 3341–3349. [Google Scholar] [CrossRef]
  49. Zhu, N.; Guo, Y.X.; Zhang, X.D.; Wang, F. The effects of vacancies defects and oxygen atoms occupation on physical properties of chromium silicide from a first-principles calculations. Solid State Commun. 2021, 340, 114535. [Google Scholar] [CrossRef]
  50. Sun, S.P.; Li, X.P.; Zhang, Y.; Wang, H.J.; Yu, Y.; Jiang, Y.; Yi, D.Q. Prediction of the mechanical properties of MoSi2 doped with Cr, Nb and W from first-principles calculations. J. Alloy Compd. 2017, 714, 459–466. [Google Scholar] [CrossRef]
  51. Deligoz, E.; Ozisik, H.; Ozisik, H.B. Calculation of the stability and mechanical and phonon properties of NbRuB, TaRuB, and NbOsB compounds. Philos. Mag. 2019, 99, 328–346. [Google Scholar] [CrossRef]
  52. Babesse, K.; Hammoutène, D.; Rodríguez-Hernández, P.; Muñoz, A.; Kassali, K.; Nedjar, R. High pressure study of structural, electronic, elastic, and vibrational properties of NaNb3O8. J. Alloys Compd. 2017, 725, 773–782. [Google Scholar] [CrossRef]
  53. Li, J.; Zhang, X.; Fang, Z.; Cao, X.; Li, Y.; Sun, C.; Chen, Z.; Yin, F. First-principles calculations to investigate electronic, elastic, and optical properties of one dimensional electride Y5Si3. Results Phys. 2021, 28, 104615. [Google Scholar] [CrossRef]
  54. Duan, Y.H.; Sun, Y.; Peng, M.J. First-principles investigations on Pb-Ba intermetallic compounds. Comput. Mater. Sci. 2014, 92, 258–266. [Google Scholar] [CrossRef]
  55. Chang, C.; Zhang, H. First-principles calculations to investigate elastic and thermodynamic properties of FeAlNixCrMn quinternary alloys. J. Mater. Res. Technol. 2022, 18, 1322–1332. [Google Scholar] [CrossRef]
  56. Bao, L.; Qu, D.; Kong, Z.; Duan, Y. Anisotropies in elastic properties and thermal conductivities of trigonal TM2C (TM = V, Nb, Ta) carbides. Solid State Sci. 2019, 98, 106027. [Google Scholar] [CrossRef]
  57. Bouchenafa, M.; Sidoumou, M.; Halit, M.; Benmakhlouf, A.; Bouhemadou, A.; Maabed, S.; Bentabet, A.; Bin-Omran, S. Theoretical investigation of the structural, elastic, electronic and optical properties of the ternary indium sulfide layered structures AInS2 (A = K, Rb and Cs). Solid State Sci. 2018, 76, 74–84. [Google Scholar] [CrossRef]
  58. Ma, G.L.; Gu, J.B.; Sun, B.; Wang, C.J.; Li, S.; Zhang, W.W. Electronic and optical properties of adsorbed of Zr-atom on vacancy-defective graphene: First principles calculations. Comput. Theor. Chem. 2021, 1206, 113468. [Google Scholar] [CrossRef]
  59. Wang, S.L.; Pan, Y. Insight into the structures, melting points, and mechanical properties of NbSi2 from first-principles calculations. J. Am. Ceram. Soc. 2019, 102, 4822–4834. [Google Scholar] [CrossRef]
  60. Chen, J.; Zhang, X.; Li, D.; Liu, C.; Ma, H.; Ying, C.; Wang, F. Insight into the vacancy effects on mechanical and electronic properties of tantalum silicide. Ceram. Int. 2020, 46, 4595–4601. [Google Scholar] [CrossRef]
  61. Han, N.N.; Liu, H.S.; Zhao, J.J. Novel magnetic monolayers of transition metal silicide. J. Supercond. Nov. Magn. 2015, 28, 1755–1758. [Google Scholar] [CrossRef]
  62. Ertürk, E.; Gürel, T. Ab initio study of structural, elastic, and vibrational properties of transition-metal disilicides NbSi2 and TaSi2 in hexagonal C40 structure. Physica B 2018, 537, 188–193. [Google Scholar] [CrossRef]
  63. Thieme, M.B.; Gemming, S. Elastic properties and electronic structure of vanadium silicides-a density functional investigation. Acta Mater. 2009, 57, 50–55. [Google Scholar] [CrossRef]
  64. Wang, S.; Pan, Y.; Wu, Y.; Lin, Y. Insight into the electronic and thermodynamic properties of NbSi2 from first-principles calculations. RSC Adv. 2018, 8, 28693–28699. [Google Scholar] [CrossRef] [PubMed]
  65. Wu, Y.; Yang, C.; Duan, Y. First-principles exploration of elastic anisotropy and thermal properties of the C40-type VSi2, NbSi2, and TaSi2 disilicides. Mater. Today Commun. 2021, 29, 102818. [Google Scholar] [CrossRef]
Figure 1. Crystal structure of C11b-VSi2 with different vacancies. (ac): side views; (d): three-dimensional view.
Figure 1. Crystal structure of C11b-VSi2 with different vacancies. (ac): side views; (d): three-dimensional view.
Metals 12 01625 g001
Figure 2. Calculated phonon dispersion curves for (a) perfect VSi2, (b) Si-va1, (c) Si-va2, (d) V-va1 and (e) V-va2, respectively. The capital letters on the horizontal axis indicate the high symmetry points in the first Brillouin zone.
Figure 2. Calculated phonon dispersion curves for (a) perfect VSi2, (b) Si-va1, (c) Si-va2, (d) V-va1 and (e) V-va2, respectively. The capital letters on the horizontal axis indicate the high symmetry points in the first Brillouin zone.
Metals 12 01625 g002
Figure 3. Surface constructions of bulk moduli of C11b-VSi2.
Figure 3. Surface constructions of bulk moduli of C11b-VSi2.
Metals 12 01625 g003
Figure 4. Surface constructions of Young’s moduli of C11b-VSi2.
Figure 4. Surface constructions of Young’s moduli of C11b-VSi2.
Metals 12 01625 g004
Figure 5. Two-dimensional projections of bulk and Young’s moduli on the (001) and (100) planes of C11b-VSi2. (a) Young’s moduli on the (001) plane; (b) Bulk moduli on the (001) plane; (c) Young’s moduli on the (100) plane; (d) Bulk moduli on the (100) plane.
Figure 5. Two-dimensional projections of bulk and Young’s moduli on the (001) and (100) planes of C11b-VSi2. (a) Young’s moduli on the (001) plane; (b) Bulk moduli on the (001) plane; (c) Young’s moduli on the (100) plane; (d) Bulk moduli on the (100) plane.
Metals 12 01625 g005
Figure 6. Electron density differences of C11b-VSi2 with various vacancies, (a) perfect C11b-VSi2, (b) Si-va1, (c) Si-va2, (d) V-va1, and (e) V-va2, respectively.
Figure 6. Electron density differences of C11b-VSi2 with various vacancies, (a) perfect C11b-VSi2, (b) Si-va1, (c) Si-va2, (d) V-va1, and (e) V-va2, respectively.
Metals 12 01625 g006
Figure 7. Total and partial density of states for C11b-VSi2 with different vacancies, (a) C11b-VSi2, (b) V-va1, (c) V-va2, (d) Si-va1, and (e) Si-va2, respectively.
Figure 7. Total and partial density of states for C11b-VSi2 with different vacancies, (a) C11b-VSi2, (b) V-va1, (c) V-va2, (d) Si-va1, and (e) Si-va2, respectively.
Metals 12 01625 g007
Table 1. The lattice parameters and vacancy formation energy Ef (eV/atom) of C11b-VSi2 with various vacancies.
Table 1. The lattice parameters and vacancy formation energy Ef (eV/atom) of C11b-VSi2 with various vacancies.
PhaseabcEfRefs.
VSi23.1783.1787.7−0.297present
-3.1553.1557.729-[10]
V-va13.1343.1337.78−0.186present
V-va23.1343.1347.779−0.185present
Si-va13.1173.1167.851−0.151present
Si-va23.1183.1187.851−0.152present
Table 2. The elastic constants Cij (GPa) of C11b-VSi2 with various vacancies.
Table 2. The elastic constants Cij (GPa) of C11b-VSi2 with various vacancies.
PhaseC11C12C13C33C44C66
VSi2198.7125.797.2390.9140.995.7
Si-va1224148.292346.4155.1145.9
Si-va2222.2149.591.2341.2152.7147.5
V-va1201159.887.2345.6161.3149
V-va2203.1155.684.8348162149.1
Table 3. Calculated elastic compliance constants Sij (GPa−1) of C11b-VSi2.
Table 3. Calculated elastic compliance constants Sij (GPa−1) of C11b-VSi2.
PhaseS11S12S13S33S44S66
VSi20.00867−0.00504−0.000910.003010.007100.01045
Si-va10.00818−0.00511−0.000820.003320.006450.00685
Si-va20.00845−0.00536−0.000820.003380.006550.00678
V-va10.01369−0.01052−0.000810.003290.006200.00671
V-va20.01220−0.00909−0.000730.003250.006170.00671
Table 4. Calculated bulk modulus, B (GPa), shear modulus, G (GPa), Young’s modulus, E (GPa) Poisson’s ratio, v and Vickers hardness, HV (GPa) of C11b-VSi2 with various vacancies.
Table 4. Calculated bulk modulus, B (GPa), shear modulus, G (GPa), Young’s modulus, E (GPa) Poisson’s ratio, v and Vickers hardness, HV (GPa) of C11b-VSi2 with various vacancies.
PhaseBVBRBHGVGRGHEνBH/GHHV
VSi2159.0150.6154.8106.682.194.3235.20.2471.64213.1
Si-va1161.8160.4161.1122.390.9106.7262.10.2291.51015.7
Si-va2161.1159.8160.4121.288.9105258.70.2311.52815.3
V-va1157.3155.7156.5121.965.793.8234.50.2501.66812.8
V-va2156.0154.3155.2122.771.196.9240.70.2411.60013.7
Table 5. Calculated elastic anisotropy indeces of C11b-VSi2.
Table 5. Calculated elastic anisotropy indeces of C11b-VSi2.
PhaseAUAcompAshearA1A2A3
VSi21.54790.02710.12981.42611.43052.6219
Si-va11.73590.00430.14731.60561.62553.8829
Si-va21.82480.00400.15371.60321.63594.0859
V-va14.28730.00510.29961.73351.72817.2068
V-va23.63970.00550.26631.69861.68946.3582
Table 6. Calculated directional elastic moduli in three principal directions (in GPa) of C11b-VSi2.
Table 6. Calculated directional elastic moduli in three principal directions (in GPa) of C11b-VSi2.
PhaseEB
[100][010][001][100][010][001]
VSi2115.3115.3332.2121.3121.3277.3
Si-va1122.2122.2301.2146.7146.7196.4
Si-va2118.3118.3295.9145.4145.4189.7
V-va173.073.0303.9139.8139.8197.6
V-va282.082.0307.7138.7138.7184.4
Table 7. The bond lengths of Si-Si covalent bonds and V-Si bonds for perfect C11b-VSi2 and C11b-VSi2 with different atomic vacancies.
Table 7. The bond lengths of Si-Si covalent bonds and V-Si bonds for perfect C11b-VSi2 and C11b-VSi2 with different atomic vacancies.
Lengths (Å)VSi2Si-va1Si-va2V-va1V-va2
Si-Si2.5292.5552.5522.4722.458
V-Si2.5782.5302.5312.5282.526
Table 8. The density ρ, sound velocities (longitudinal νl, transverse νt and mean νm) and Debye temperature θD.
Table 8. The density ρ, sound velocities (longitudinal νl, transverse νt and mean νm) and Debye temperature θD.
Phaseρ (g/cm3)νl (m/s)νt (m/s)νm (m/s)θD (K)
VSi24.5617842.44546.83570.6452
V-va14.3798018.84628.33635.1457
V-va24.3788061.24706.83695.3465
Si-va14.5138198.94862.53815.6480
Si-va24.5098161.84825.33786.9476
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Xu, S.; Duan, Y.; Peng, M.; Shen, L. Effects of Vacancies on the Structural, Elastic, Electronic and Thermodynamic Properties of C11b-VSi2 by First-Principles Calculations. Metals 2022, 12, 1625. https://doi.org/10.3390/met12101625

AMA Style

Xu S, Duan Y, Peng M, Shen L. Effects of Vacancies on the Structural, Elastic, Electronic and Thermodynamic Properties of C11b-VSi2 by First-Principles Calculations. Metals. 2022; 12(10):1625. https://doi.org/10.3390/met12101625

Chicago/Turabian Style

Xu, Shan, Yonghua Duan, Mingjun Peng, and Li Shen. 2022. "Effects of Vacancies on the Structural, Elastic, Electronic and Thermodynamic Properties of C11b-VSi2 by First-Principles Calculations" Metals 12, no. 10: 1625. https://doi.org/10.3390/met12101625

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