Next Article in Journal
Hot Deformation Behavior of Fe40Mn20Cr20Ni20 Medium-Entropy Alloy
Previous Article in Journal
Surface Characterization of Carbon Steel after Rolling Burnishing Treatment
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Effect of Point Defects on Young’s Modulus of the Off-Stoichiometric Al3X (X = Li, Sc, and Zr) Phases: A First-Principles Study

1
School of Materials Science and Engineering, Beijing Institute of Technology, Beijing 100081, China
2
Advanced Research Institute of Multidisciplinary Science, Beijing Institute of Technology, Beijing 100081, China
*
Author to whom correspondence should be addressed.
Metals 2024, 14(1), 30; https://doi.org/10.3390/met14010030
Submission received: 4 November 2023 / Revised: 2 December 2023 / Accepted: 7 December 2023 / Published: 26 December 2023
(This article belongs to the Section Metal Casting, Forming and Heat Treatment)

Abstract

:
L12-Al3X (X = Li, Sc, and Zr) precipitates are the main strengthened phases of high-strength aluminum alloys and are critical for aerospace structural materials. Point defects and substitutional ternary elements change the mechanical properties of Al3X. In this paper, the effect of point defects, including vacancy, antisite, and substitutional element addition defects on the elastic modulus of the off-stoichiometric Al3X (X = Li, Sc, and Zr) phase were investigated by using first-principle calculations. The formation enthalpies of the defective Al3X alloy and isolated point defects in Al3X were calculated, and the results showed that the defects have an effect on the structure and elasticity of the off-stoichiometric Al3X phases. The lattice distortion, elastic constants, and elastic moduli were further investigated. It was found that the point defects increased the Young’s modulus for Al3Zr, and the doping of Er improved the Young’s modulus for off-stoichiometric Al3Li and Al3Sc. Adjusting the position of vacancies can improve the elastic modulus. In addition, the doping of substitutional elements (especially Sc, Ti, Zr, Hf, Ta, Mn, Ir, and Cf) can greatly increase the Young’s modulus of off-stoichiometric Al3Li.

1. Introduction

L12-Al3X (X = Li, Sc, and Zr) precipitates in aluminum alloys are beneficial for improving elastic properties [1,2,3,4,5]. Many studies have shown that defects and foreign element doping play an important role in elastic properties [6,7,8,9,10]. Although many point defects in Al3X have been reported in experiments by researchers, the underlying kinetic transformation process is still unclear, primarily due to the difficulties in imaging the vacancies. Moreover, it is important to understand the effect of different point defect types on the elastic properties of structural materials.
Recently, based on Density Functional Theory (DFT), the first-principle calculations have provided another dimension of investigating the point defects computationally. For example, Dong et al. [11] found that point defects positively affect the brittle behavior of Hf5BSi3. Jiang et al. [12] investigated the effect of point defect on the structures and mechanical properties of defective BiCuOTe, and they found that point defects improved the elastic compliance but weakened the compression resistance of BiCuOTe. Fan et al. [13] suggested that Al matrix with vacancy and substitutional rare element addition presented the ductility and nonuniform. Muzy et al. [14] studied the point defects by thermostimulated luminescence, and it was found that the dislocations in the specimens were pinned due to point defects, while the samples with more dislocations became more resistant to plastic deformation in Verneuil crystals. Ouahrani et al. [15] studied the effect the point defects on Cu2WS4 single layer and found that the vacancy changed to the energy structure. Wang et al. [16] found that point defects cause lattice distortion for Al3Sc and Al3Lu, which reduces the elastic properties. In addition, the microalloying of rare earth elements has been recently reported to have an impact on the precipitation structure and, thus, its elastic properties [17,18,19,20,21,22]. Zhang et al. [17] proposed that the addition of Sc to Al-Cu alloys promoted the nucleation of Al grains by formation Al3Sc, reduced the average size of θ′ precipitates, and increased the number density of θ′ precipitates, thereby improving the yield strength and tensile strength of Al alloys. Ying et al. [8] and Wang et al. [9] also suggested that the addition of Sc improved the grain refinement and thus improved the mechanical properties of the Al alloy. Moreover, it was found that the addition of Sc, Zr to aluminum alloys produced the Al3(Sc, Zr) particles with core-shell structure, which refined the grain size [2], and improved the tensile strengths [3] and yield strengths [18,19]. Yi et al. [23] found that the doping of Sc had positive effects on the enhancement of creep resistance in the Al-Ce alloy. Ce promoted the formation of second phases and grain refinement [24,25], and Ekaputra et al. [20] found that the combination of the Al11Ce3 and L12-Al3(Er, Sc, Zr) precipitates increased the strength of the aluminum alloy. It was found that the addition of Er to the Al-Sc alloys promoted the formation of an Al3(Er, Sc) structure, improving the elastic properties of Al alloys [21,22]. Fang et al. [26,27,28] revealed that Yb additions could enhance the fracture toughness of Al-Zn-Mg-Cu alloy, and they also found that Pr could improve the strength, fracture toughness, and ductility of Al-Zn-Mg-Cu-Zr alloy [29]. Tian et al. [30] investigated the effects of Sc on the elastic properties of Al3Li by first-principles calculations and found that the doping of Sc enhanced the shear deformation resistance and stiffness of Al3Li. Compared with rare earth elements, generally, Er exhibits a relatively large diffusivity in α-Al [31] and thus shows a robust defect formation activity. Notably, full information on the influences of Er on the elastic properties is of great value for the development and design of materials. There are a large number of vacancies in Al3X in Al alloys, and previous experiments reported that the doping of Er led to the formation of an Al3(X, Er) structure [20,21,22,32,33]. However, the intermediate steps of how the solute atom Er substitutes the vacancy in Al3X to form a new structure and how they change the modulus in this paper are not systematically studied. The formation of the intermediate states as precipitation is experimentally difficult to observe due to the small amounts, while the first-principle calculations can analyze the structural properties of energy [34].
In general, the point defects of intermetallic compounds are complex, and the various types of vacancy, antisite, and substitute point defects co-exist, causing them to deviate from their stoichiometric ratios. The point defect with off-stoichiometric ratios both have physical significance in Al3X phases. In this work, we investigated the effects of point defects and the alloying element Er on the structural stability and elastic properties of off-stoichiometric Al3X while considering the new structures as pure precipitations by using first-principle calculations. The effects of complex defect forms, such as double, triple, quadruple, and septuplet defects, are left for further study. Although we just studied the effect of only one Er atom on Al3X, knowledge of the first step of Er addition on the Al3X is of great importance for understanding the nucleation of Al3(X, Er), which may be useful in the design of guided experiments for the development of new materials. To prove the reliability of our model structures, we measured Young’s modulus (E) of Al3Sc particles by a nanoindentation experiment. The occupation behavior and the elastic influence of the Er-added point defects in Al3X were further analyzed. This lays a foundation for the application of point defects and Er in aluminum alloys and could provide theoretical data and directions for the development of high-performance Al alloys.

2. Computational Methods

In this paper, first-principle calculations based on the Density Functional Theory (DFT) [35,36] of the effects of point defects and the Er addition on structural stability and elastic properties of off-stoichiometric Al3X (X = Li, Sc, and Zr) were performed by using the Vienna Ab-initio Simulation Package (VASP) [37,38,39,40,41]. In this study, the projector-augmented wave (PAW) [42] method was used to deal with the electron-core interactions. For the electron exchange-correlation, the Perdew–Burke–Ernzerh (PBE) functional [43] based on generalized gradient approximation (GGA) [44] was employed throughout the predictions. The Perdew–Wang parameterization (PW91) [45,46] and the local density approximation (LDA) [47,48] were only used to reproduce the previous reports and have been found to be less accurate than PBE. The plane wave cutoff energy was set to 450 eV, and the energy and force convergence criteria were less than 1.0 × 10−6 eV and 1.0 × 10−2 eV/Å, respectively. In total, there are 32 atoms in a 2 × 2 × 2 supercell for the calculation of defective structures and doped structures. The K-points were selected to be 9 × 9 × 9 for 32-atom defect phases in order to sample the Brillouin region. The optimized cutoff energy and K-points were validated by taking the residual energy less than 3 × 10−3 eV.
To check the effect of supercell size, we took the unit cell of the Al3X phase as an example to perform a convergence test by considering the supercell of 2 × 2 × 2 (32 atoms), 3 × 3 × 3 (108 atoms), and 4 × 4 × 4 (256 atoms). The formation enthalpies for these different supercells were calculated, and the difference among them is less than 0.01 eV/atom. Therefore, a 2 × 2 × 2 supercell was used for calculation. We have performed the total energies and the formation enthalpies of the optimized structures in Table S1. The formation enthalpies for pure Al3X supercell and defected Al3X structure with point defect could be calculated as [49,50]
H = ( E t o t a E A l b E X ) / ( a + b )
where E t o t is the total energy of the pure or defective off-stoichiometric Al3X, and X represents the atoms of Li, Sc, or Zr. E A l and E X represent the energy of each element at the bulk states of Al and X bulk materials, respectively. a and b are the atom numbers of Al and X in the pure or defective Al3X, respectively. According to the Wagner-Schottky model, the formation enthalpy as a function has a linear relationship with the point defect concentrations,
H = H A l 24 X 8 + d H d x d
where x d is the atomic concentration in point defects. H A l 24 X 8 is the formation enthalpy (∆H) of the full-ordered stoichiometric Al24X8 supercell, and the subscript d represents the point defect types. H d is the formation enthalpy of the point defect in stoichiometric Al3X. The point defect formation enthalpy (Hd) was also calculated by finite differencing [51,52],
H d = Δ H x d H d H A l 24 X 8 x d
x d = 1 / ( a + b )
where H d is the formation enthalpy of defective Al3X structures. The phase stability of off-stoichiometric Al3X with doping Er could be evaluated by the formation enthalpy (ΔH),
H = ( E t o t a E A l b E X E E r ) / ( a + b + 1 )
where E t o t is the total energy of the doped off-stoichiometric Al3X, and E E r represent the energy of the Er element at the bulk state. a and b are the atom numbers of Al and X in the doped off-stoichiometric Al3X.
The elastic constants (Cij) of the crystal are a measure of the resistance to external pressure, playing an important role in mechanical and dynamical behaviors. In this work, the elastic constants (Cij) were obtained by the coefficient of the stress-strain curve [53]. There are two methods of calculating elastic constants from the first principles: the stress–strain and the energy–strain. In the condition of small strains, there was a linear relationship satisfying Hooke’s law between strain ε = (ε1, ε2, ε3, ε4, ε5, ε6) and stress σ = (σ1, σ2, σ3, σ4, σ5, σ6):
σ i = j = 1 6 C i j ε j
where σ i and ε j represented strain and stress with six independent components respectively. Cij represented the elastic constant with subscripts using the Voigt notation. For the stress–strain method, based on the first-order derivative of the stresses, the corresponding stresses are calculated in terms of different strains. The primary term coefficients can be obtained by fitting through Equation (5) to obtain the elastic constants Cij. There is an energy variation of the equilibrium lattice structure under applied minor strains. In a second-order approximation, the elastic energy E V , ε i of a solid under strain can be given by Taylor expansion:
E V , ε i = E V , ε i E ( V 0 , 0 ) = V 0 2 j = 1 6 C i j ε i ε j
where E V , ε i and E ( V 0 , 0 ) are the total energy of the structure with and without applied strain, respectively, and V 0 is the equilibrium volume. For the energy–strain method, the energy variation of the solid under the different strains is calculated, and the elastic constants are derived from the second-order derivatives of the energy versus strain curves by fitting Equation (6). For each strain, there are six independent stress components but only one energy to fit, obtaining the elastic constants. Compared to the energy–strain method, the stress–strain method requires a smaller set of distortions and is more efficient [54,55,56,57]. In this paper, the calculations of the elastic constants have been performed using the stress–strain approach. When a train ε is applied to deform the crystal, the lattice vectors between the distorted and equilibrium cells can be expressed as:
a b c = a b c · ( I + ϵ )
where I is the unit matrix. The strain tensor ϵ and elastic constants C i j of the fcc structure Al3X can be described by
ϵ = e 1 e 6 2 e 5 2 e 6 2 e 2 e 4 2 e 5 2 e 4 2 e 3 2
C i j = C 11 C 12 C 12 C 12 C 12 0 C 11 C 12 0 C 11 C 11 0 0 0 0 0 0 0           0 0 0 0 0 0 0 0 0 C 44 0 0 0 0 C 44 0 0 C 44
then substituting the Equation (9) into Equation (5), the three-independent elastic constants (C11, C12, C44) of cubic crystal can be calculated by
σ 1 σ 2 σ 3 σ 4 σ 5 σ 6 = C 11 C 12 C 12 C 12 C 12 0 C 11 C 12 0 C 11 C 11 0 0 0 0 0 0 0           0 0 0 0 0 0 0 0 0 C 44 0 0 0 0 C 44 0 0 C 44 ε 1 ε 2 ε 3 ε 4 ε 5 ε 6
All crystal structures need to satisfy the mechanical criterion in this work. And the criterion to evaluate the mechanical stability of pure, defective, and doped Al3X could be expressed as [58,59,60]:
C 11 C 12 > 0 , C 11 > 0 ,   C 44 > 0 , C 11 + 2 C 12 > 0
According to the Voigt–Reuss–Hill (VRH) approach, the bulk modulus B , shear modulus G , Young’s modulus E , and Poisson’s ratios v of cubic structures are written as [61]:
E = 9 GB / ( 3 B + G ) , v = ( 3 B 2 G ) / 2 ( 3 B + G )
B V = B R = ( C 11 + 2 C 12 ) / 3
G = ( G V + G R ) / 2
G V = C 11 C 12 + 3 C 44 / 5
G R = 5 ( C 11 C 12 ) C 44 / [ 4 C 44 + 3 ( C 11 C 12 ) ]
where, the subscripts V and R represent Voigt and Reuss approximations, respectively.
The number of independent elastic constants depends on the symmetry of crystals. The lower the symmetry, the more the independent elastic constants. For example, there are three independent elastic constants for Al3X cubic crystals, namely C11, C12, and C44 [62,63]. For the defective off-stoichiometric Al3X, the point defect changes the composition of the chemical formula and the symmetry of the crystals, and the ratio of Al to X is no longer 3:1 (as shown in Table S2). These structures, without the L12 structure, will have more independent elastic constants, so the elastic moduli were calculated with the help of the VASPKIT program [64]. According to the coefficient of the stress–strain curve, VASPKIT can identify the symmetry of the defective supercell structure and obtain the corresponding stiffness tensor matrix, acquiring independent elastic constants as well as polycrystalline elastic modulus.

3. Experimental Procedure

To verify our theoretical calculation, we investigated the commercial Al-2Sc binary alloy (as cast) with a 2% mass fraction of Sc. The Al-2Sc cast has a large size of primary Al3Sc particles at a low solidification rate [65]. Using a continuous stiffness model [66], Young’s modulus E of Al3Sc particles was measured by nanoindentation experiments. All measurements were done on the Keysight G200 Nanoindenter, with a diamond Berkovich-type tip at a room temperature of 23.7 °C. In this work, the simulation calculations were performed at T = 0 k, and the room temperature in the experiment can correspond with our calculation conditions to verify our calculation results. Fused quartz was chosen as a reference sample for calibration of the tip area function. To keep the reliability of the experimental results, the test selected peak loads ranges from 1 to 10 mN, and the indentations were made on at least five different Al-2Sc particles for each peak load, and the nanoindentation depth stabilized between 50 nm and 150 nm while the Al3Sc particle size was intentionally annealed to the size over 150 nm. Data were analyzed by using NanoSuite 6.40.458 software (Copyright 2015 Keysight Technologies, Santa Rosa, CA, USA) to determine the starting position of the indenter contacting the specimen surface to calibrate the load indentation curve. The modulus of elasticity was measured from the contact area versus the continuous stiffness.

4. Results and Discussion

In this paper, the structural stability and elastic properties of defective and Er-doping off-stoichiometric Al3X phases were calculated. First, the bulk properties (lattice constants and bulk modulus) of L12-Al3X were calculated. Then, a nanoindentation experiment on the primary Al3X phase was performed. Finally, the stability and elastic properties of defective and doped off-stoichiometric Al3X were obtained, clarifying the effects of the point defects and Er additions.

4.1. The Bulk Properties

The L12-Al3X precipitates are face-centered cubic crystal structures (see Figure 1) [5,67,68,69,70,71]. To study the ground-state properties of Al3Li, Al3Sc, and Al3Zr, their equilibrium lattice configurations using different methods for the electron exchange-correlation functionals were calculated and compared to previous experiments [72,73,74,75,76,77,78]. The calculated ground-state properties are shown in Table 1. The lattice constants (a0) and bulk modulus (B0) agree well with the experimental data and prior calculations [75,76,77,78,79,80,81,82]. There were no experimental data of a0 and B0 for Al3Zr, so the comparisons of Al3Li and Al3Sc to experiments were performed. It is clearly shown that our calculations using PBE have a better match with the results of previous experiments compared to those using LDA or PW91. Although the calculation overestimates (underestimates) the a0 (B0) of Al3Li by 0.4% (3%) and Al3Sc by 0.05% (3%), the accuracy is within the reliability range of computational models [76,78,79,80,83,84].
Before we discuss the optimal doping position of Er in off-stoichiometric Al3X, it is necessary to have a good understanding of the nature of point defects in Al3X. The point defects lead to small deviations in the stoichiometric ratios of the chemical compositions with respect to the fully ordered L12-Al3X, formatting the Al-rich and X-rich off-stoichiometric Al3X. For the 32-atom Al3X, there are four possible types of point defects, as shown in Figure 2(a1–a4), designated as VAl, XAl, VX, and AlX, respectively. Considering the large number of free Al atoms from the aluminum matrix filling around the Al3X crystals, the different types of point defects formed were discussed. For Al-rich off-stoichiometric Al3X, the X atom in Al3X migrates and leaves VX. The vacancy would be occupied by the Al atom from inside the Al3X or Al matrix, forming AlX. For X-rich off-stoichiometric Al3X, the Al atom migrates and reforms the vacancy of VAl; the X atom from Al3X could occupy the vacancy to form an antisite XAl. The atoms were allowed to occupy their ideal lattice positions in the off-stoichiometric Al3X supercell at the beginning, and then local relaxation of atoms around the point defect was performed in order to enable all atoms to move to their equilibrium positions. To better understand the lattice structure, the mean distance of n-th nearest neighbors (nNN) of atomic pairs in Al3X were counted and shown in Figure 2b. The point defects caused the displacement of atoms, and the specific data is shown in Table S3.
In general, the lower the alloy formation enthalpy of defective Al-rich or X-rich off-stoichiometric Al3X supercells are, the better their stability would be [49,85]. Fully relaxed formation enthalpies of ordered stoichiometric Al3X, Al-rich, and X-rich off-stoichiometric Al3X were listed in Table 2 and Figure 3. It can be seen that the formation enthalpies of defective off-stoichiometric Al3X supercell alloys were negative, indicating their thermodynamic stability [86]. The results of Al3Li, Al3Sc, and Al3Zr in Table 2 were also consistent with previous theoretical calculations [16,62,86,87]. For X-rich off-stoichiometric Al3X alloys, the formation enthalpies of Al antisite (XAl) (Al3Li—0.101 eV/atom, Al3Sc—0.435 eV/atom, and Al3Zr—0.481 eV/atom) are smaller than that of Al vacancy, meaning that Al antisite defect (XAl) are stable. In addition, the results showed that the stable point defects in L12 Al3X are antisite on both Al-rich and X-rich off-stoichiometry. For off-stoichiometric Al3Sc and Al3Zr, the results showed that the stable point defects are Al antisites in X-rich and X vacancies in X-rich off-stoichiometry.
The point defect changed the distance of the surrounding atoms. To quantify and distinguish the structural properties of the defective off-stoichiometric Al3X, we calculated radial distribution functions (RDFs) of the Al-Al, X-X, and Al-X atomic pairs within these defective structures and took a comparison. The results are shown in Figure 4(1a–3d). The RDF as a key quantity for evaluating the local distribution of atomic pairs can be expressed as [88]:
g r = n r ρ 0 V n r 4 π r 2 ρ 0 δ r
where r is the cutoff radii from the center atoms, δ r is the spherical shell thickness,   n r is the number of particles inside the spherical shell, and ρ 0 is the median density of the ideal crystal. As seen in Figure 4(1a–3d), there is a sharp peak near the first nearest neighbor (1NN) for all defective structures, which indicates the density of the first nearest neighbor atomic pairs. The low peaks in the third and fifth nearest neighbors indicate a weak intensity. The calculation results are very similar for the vacancy (VAl and VX), both having two peaks in the 1NN–5NN. For the XAl antisite, there are three peaks in the 1NN and 3NN. In brief, there are relatively weak changes in atom displacements relative to the ideal position.
The site preference of Er in off-stoichiometric Al3X was calculated by taking four occupation sites in off-stoichiometric Al3X: (i) the Er atom directly occupying the Al site (ErAl); (ii) the Er atom replacing the Al site meanwhile the Al atom occupying X site (ErAl + AlX); (iii) the Er directly taking up the X site (ErX); and (iv) the Er replacing the X site meanwhile X atom occupying the Al site (ErX + XAl). The four occupation positions of Er occupying 32-atom Al3X are shown in Figure 5a–d.
The site preference of the Er element in off-stoichiometric Al3X was evaluated by calculating the enthalpy at absolute temperature. According to Equation (5), the enthalpies were listed in Table 3, and all doped Al3X phases were stable due to their negative formation enthalpies. For X-rich off-stoichiometric Al3Li, it was found that the formation enthalpy of off-stoichiometric Al3Li (ErLi + LiAl) (−0.189 eV/atom) is lower than off-stoichiometric Al3Li (ErAl) (−0.125 eV/atom). For Al-rich off-stoichiometric Al3Li, the formation enthalpy of Al24Li7Er (ErLi) (−0.171 eV/atom) is lower than Al24Li7Er (ErAl + AlLi) (−0.112 eV/atom), revealing that the Er atom prefers to substitute the Li atom. Meanwhile, improving the point defect structure of Er indirectly occupying the Li site is more stable compared to direct occupancy. As a result, the lattice parameter is 4.059 Å. For off-stoichiometric Al3Sc and Al3Zr, the formation enthalpies show that the Er atom prefers to occupy the Sc (Zr) site in Al-rich and Sc (Zr)-rich off-stoichiometric Al3Sc (Al3Zr) supercells directly, increasing the lattice parameter to 4.123 Å (4.119 Å). The RDF diagram showed the change in crystal structure after Er doping (as shown in Figure 6). For the Er-Al and Er-Li bonds, there are sharp peaks in 1NN, 2NN, 3NN, and 4NN. In particular, there are additional Er-Al peaks at 1NN and 3NN and additional Er-Li peaks at 2NN and 4NN for the ErX + XAl, indicating the doping of Er changed the structure and increased the combination of bonds. The atom radius of Er (1.75 Å) [89] is similar to that of Sc (1.57 Å) [90] and Zr (1.60 Å) [91], while the atomic radius of Al (1.43 Å) is close to that of Li (1.52 Å) [92]. Therefore, Er is likely to occupy the Sc (Zr) sites, while Al prefers to occupy the Li sites. The doping of the Er element induces a local stress field and tends to replace the similar-sized atoms [93], which is the essential condition for the stabilization of the doped phases. Therefore, Er prefers to occupy the X atom site, not to occupy the Al atom site.

4.2. Elastic Modulus

The elastic properties of materials reflect their strain resistance, which plays the role of one of the most important mechanical properties [94,95,96]. In fact, these properties are very important for the industrial applications of engineering materials. To study the effect of the point defects and Er doping on the elastic properties of these phases, the elastic constants (Cij), elastic modulus (B, G, and E), υ, and B/G of pure, defective, and doped off-stoichiometric Al3X at 0 K were calculated. The values of modulus reflect the bonding strength between atoms in crystals. The ratio of B/G indicates the size of the plasticity range, and a high value is associated with malleability and a low value with brittleness [62,97]. Meanwhile, the corresponding previous experimental and calculated data in the literature [30,76,78,79,80,81,82,98,99,100] of pure Al3X were used to verify the accuracy of our calculations, as shown in Table 4 and Table S1.
As listed in Table 4, the calculated results of bulk modulus B, shear modulus G, Young’s modulus E, Poisson’s ratios v, and ratios of the bulk modulus to shear modulus B/G for the cubic phase Al3Li are 63.8 GPa, 42.7 GPa, 104.7 GPa, 0.226, and 1.495, respectively, which are close to the experimental results [79] (66.0 GPa, 43.0 GPa, 105.9 GPa, 0.23, and 1.535, respectively). Additionally, our calculation values of Al3Li are nearly the same as previous calculations [30,76,98], as shown in the Table S4. For the pure Al3Sc phase, our calculation data of bulk modulus B, shear modulus G, Young’s modulus E, Poisson’s ratios v, and ratios of the bulk modulus to shear modulus B/G are 88.2 GPa, 69.4 GPa, 164.9 GPa, 0.189, and 1.272, respectively, which agrees well with the experimental data [78] and calculated data [76,78,80,99,100]. To further ensure the reliability of the calculated data, as described in Section 3, we measured and calculated Young’s modulus E by nanoindentation experiments. Overall, thirteen effective curves of Young’s modulus and indentation depth were obtained by measuring different Al-2Sc particles. As shown in Figure 7, the E of Al3Sc was determined as 170.1 ± 3.9 (GPA), which was in good agreement with our calculated value (171.1 GPA) and Woodward’s value (164.9 GPa). Similarly, the calculated results of elastic properties of pure Al3Zr were close to previous reports [80,81,82]. Those data further confirmed the validity of our model and the reliability of our calculation results.
To study the influence of defect points and Er doping, we further calculated and compared the elastic properties of pure, defective, and doped Al3X (X = Li, Sc, and Zr). The defective and doped Al3X all satisfied the mechanical stability according to Equation (12). The results showed that the defective Al3X is not only energetically stable but also mechanically stable. As seen in Table 4, combining the elastic modulus (B, G, and E), Poisson’s ratios, and ratios of the bulk modulus to shear modulus of pure and defective off-stoichiometric Al3X, we could find that: (i) compared with other types of defects (VAl, LiAl, and AlLi), the VLi defect could improve the elastic modulus (B, G, E) while maintaining Poisson’s ratios, and ratios of the bulk modulus to shear modulus similar to pure Al3Li. Since LiAl in Li-rich and AlLi in Al-rich off-stoichiometric Al3Li were relatively stable, we speculated that the existence of point defects would increase the bulk modulus, Poisson’s ratios, and ratios of the bulk modulus to shear modulus of off-stoichiometric Al3Li. Those results mean that defective Al3Li has a higher resistance to deformation and ductility than pure Al3Li [94,101]; (ii) the existence of point defect (except for VSc) increased Poisson’s ratios and ratios of the bulk modulus to the shear modulus of off-stoichiometric Al3X, meanwhile decreasing the elastic modulus (except for bulk modulus of VAl). Because the ScAl in Sc-rich and VSc in Al-rich off-stoichiometric Al3Sc were stable, we predicted that the point defects would reduce the bulk modulus, shear modulus, and Young’s modulus. The results showed that point defects decrease the resistance to the strain of Al3Sc crystal, making it easy to undergo deformation; and (iii) the presence of VAl and ZrAl increases the elastic modulus and Poisson’s ratios and ratios of the bulk modulus to the shear modulus of off-stoichiometric Al3Zr. AlZr decreased its elastic modulus while increasing Poisson’s ratios and ratios of the bulk modulus to the shear modulus. Additionally, VZr decreased its Poisson’s ratios, ratios of the bulk modulus to shear modulus, and its bulk modulus. Because ZrAl in X-rich and VSc in Al-rich off-stoichiometric Al3Zr were stable, we inferred that the point defects would enhance the Young’s modulus of off-stoichiometric Al3Zr. The results suggested that defective Al3Zr were less prone to deformation than pure stoichiometric Al3Zr. The lattice parameter of Al bulk is 4.038 Å. The lattice mismatch of defective Al3X with Al bulk became larger, and that of defective off-stoichiometric Al3Zr with Al bulk became small, which is inversely related to the elastic modulus. Compared the elastic properties of pure Al3X with that of Er-doping off-stoichiometric Al3X, we could conclude that: (i) the substitution of Al to the Er site led to the decrease of elastic modulus while increasing Poisson’s ratios, and ratios of the bulk modulus to shear modulus of off-stoichiometric Al3Li. The substitution of Li to Er and Al to Li sites increased the shear modulus, Young’s modulus, Poisson’s ratios, and ratios of the bulk modulus to the shear modulus of off-stoichiometric Al3Li. The substitution of Li to Er site increases the elastic modulus (B, G, E), Poisson’s ratios, and ratios of the bulk modulus to shear modulus of off-stoichiometric Al3Li. The substitution of Al to Er and Li to Al sites increases the bulk modulus, Poisson’s ratios, and ratios of the bulk modulus to shear modulus of off-stoichiometric Al3Li. Because the Er atom preferred to occupy the ErLi + LiAl site, we could conclude that the doping of Er increased the shear modulus and Young’s modulus of off-stoichiometric Al3Li. The results indicated that the Al3Li with doping Er has a higher resistance to deformation and that of shear strain; (ii) the substitution of Al to Er site increases the bulk modulus, Poisson’s ratios, and ratios of the bulk modulus to the shear modulus of off-stoichiometric Al3Sc. The substitution of Sc to Er and Al to Sc sites and Sc to Er sites both increase the elastic modulus of Al3Sc, and the increment of occupying ErSc site was less than that of the ErSc + ScAl site relatively. ErAl + AlSc decreases the elastic modulus, Poisson’s ratios, and ratios of the bulk modulus to the shear modulus. Because of the preference of the ErSc, we could find that the doping of the Er atom increased the bulk modulus, shear modulus, Young’s modulus, and ratios of the bulk modulus to the shear modulus while decreasing Poisson’s ratios of off-stoichiometric Al3Sc. The results demonstrated that Er-doping improved resistance to strain and deformation while increasing plasticity; and (iii) the substitution position of ErAl, ErZr + ZrAl, ErZr, and ErAl + AlZr all decreased the elastic modulus of off-stoichiometric Al3Zr. The Er atom preferred to occupy the ErZr, so we could conclude that the doping of the Er atom decreased the elastic modulus, Poisson’s ratios, and ratios of the bulk modulus to the shear modulus of off-stoichiometric Al3Zr. The results suggested that Er-doping improved the resistance to deformation and plasticity. The lattice mismatch of Er-doping off-stoichiometric Al3X with Al bulk all became larger, indicating that Er has some impact on lattice mismatch and elastic modulus of off-stoichiometric Al3X.
Combined with the relatively more stable structures of defective and Er-doping off-stoichiometric Al3X, a more intuitive variation of the pure, defective, and Er-doping elastic properties is shown in Figure 8a–e. Compared with the clean Al3Li, the bulk modulus of defective off-stoichiometric Al3Li generally increases by 5.64%. The shear modulus and Young’s modulus of defective off-stoichiometric Al3Li generally decrease by 5.62% and 3.72%, respectively. Poisson’s ratios and ratios of the bulk modulus to the shear modulus of defective off-stoichiometric Al3Li generally increase by 11.1% and 11.91%, respectively. Those results meant that defective off-stoichiometric Al3Li have smaller shear modulus and Young’s modulus but higher bulk modulus than pure Al3Li. However, the bulk modulus of Er-doping off-stoichiometric Al3Li decreases by 0.94%. Shear modulus and Young’s modulus of Er-doping off-stoichiometric Al3Li increase by 4.22% and 3.25%, respectively. Poisson’s ratios and ratios of the bulk modulus to shear modulus of Er-doping off-stoichiometric Al3Li decrease by 4.87% and 4.95%, respectively. Compared with the clean Al3Sc, the bulk modulus, shear modulus, and Young’s modulus of defective off-stoichiometric Al3Sc decrease 0.34%, 3.46%, and 2.85% respectively, and that of Er-doping off-stoichiometric Al3Sc increase 5.67%, 2.45%, and 3.15% respectively; Poisson’s ratios and ratios of the bulk modulus to shear modulus of defective off-stoichiometric Al3Sc increased by 3.74% and 3.14%, and that of Er-doping Al3Sc increased by 3.64% and 2.99%. Those results indicated that defective off-stoichiometric Al3Sc have lower elastic modulus but higher Poisson’s ratios and ratios of the bulk modulus to shear modulus than pure Al3Sc, whereas the Er-doping off-stoichiometric Al3Sc have higher elastic modulus and Poisson’s ratios and ratios of the bulk modulus to shear modulus. Compared with the Al3Zr, the bulk modulus and Young’s modulus of defective off-stoichiometric Al3Zr increased by 19.99% and 1.97%, and that of Er-doping off-stoichiometric Al3Zr decreased by 2.62% and 1.29%; shear modulus of defective and Er-doping off-stoichiometric Al3Zr decreased by 12.22% and 1.06% respectively; Poisson’s ratios of defective and Er-doping off-stoichiometric Al3Zr increased by 16.46% and 21.25%; the ratios of the bulk modulus to shear modulus of defective and Er-doping off-stoichiometric Al3Zr decreased by 1.69% and 1.66%. Those results meant that the defective off-stoichiometric Al3Zr has a higher bulk modulus and Young’s modulus than pure Al3Zr, whereas Er-doping off-stoichiometric Al3Zr has a higher shear modulus.
We can increase the elastic modulus of the phase by adjusting the point defects. The effect of doping other solid solution atoms on the Young’s modulus remains to be further investigated, which is significant for screening the design of new materials with high modulus. For off-stoichiometric Al3Li, the low of Young’s modulus with solid solution atoms was summarized, as shown in Figure 9. The results show that the Young’s modulus of off-stoichiometric Al3Li tends to increase and then decrease with increasing atomic number at the same period. In particular, the atoms of elements, such as Sc, Ti, Zr, Hf, Ta, Mn, Ir, or Cf, can greatly increase the Young’s modulus.

5. Conclusions

In this paper, the point defects and the substitutional element additions to the off-stoichiometric Al3X structure were investigated by using first-principle calculations. The formation enthalpies of defective and doped off-stoichiometric Al3X were calculated, and the elastic properties of point defects and Er were analyzed. We found that the negative formation enthalpies of defective and doped off-stoichiometric Al3X indicate thermodynamic stability. By comparing the defective structure with the pure Al3X structures, the following conclusions can be drawn: (i) XAl is easier to form in off-stoichiometric Al3X, Er tends to occupy the Li site in off-stoichiometric Al3Li (ErLi + LiAl) indirectly, and Er prefers to occupy the X sites; (ii) the existence of point defects would reduce Young’s modulus of off-stoichiometric Al3Li and Al3Sc, while they can increase Young’s modulus of the off-stoichiometric Al3Zr phase; (iii) the substitution of the Er atom in the off-stoichiometric Al3Sc phase could increase elastic modulus. These results suggest that point defects and micro-alloying Er could improve the elastic properties of the Al-Li alloys; (iv) for off-stoichiometric Al3Li, the Young’s modulus of Al3Li was improved by adding such substitutional atoms as Sc, Ti, Zr, Hf, Ta, Mn, Ir, or Cf.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/met14010030/s1. Table S1. The optimized total energies and formation energies of different defective supercells. Table S2. The symmetry and lattice parameter of Al3X and the ratio of Al to X in the defective structures. Table S3. The mean displacement of the N-th neighbor in the defective Al3X. Table S4. The calculated elastic constants Cij of defective off-stoichiometric Al3X.

Author Contributions

Conceptualization, J.W. and Y.M.; methodology, Y.M.; software, Y.M.; validation, S.W., C.Z., B.W. and C.X.; formal analysis, Y.M., S.W., H.S., C.Z., B.W. and C.X.; investigation, Y.M.; resources, B.W. and C.X.; data curation, Y.M. and S.W.; writing—original draft preparation, Y.M.; writing—review and editing, J.W.; visualization, C.Z.; supervision, J.W.; project administration, J.W.; funding acquisition, J.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (grant number 52073030).

Data Availability Statement

The data presented in this study are available on reasonable request from the corresponding author. The data are not publicly available due to privacy.

Acknowledgments

The simulations were performed using resources provided by the Integrated Computing Materials Engineering (ICME) laboratory at Beijing Institute of Technology.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Liu, Z.Q.; Ma, P.P.; Jiang, Y.; Wang, Y.R.; Liu, C.H. First-principles insights into complex interplays among nano-phases in an Al-Cu-Li-Zr alloy. Acta Mater. 2022, 239, 118304. [Google Scholar] [CrossRef]
  2. Ye, J.; Pan, Q.L.; Liu, B.; Hu, Q.; Qu, L.F.; Wang, W.Y.; Wang, X.D. Effects of co-addition of minor Sc and Zr on aging precipitates and mechanical properties of Al-Zn-Mg-Cu alloys. J. Mater. Res. Technol. 2023, 22, 2944–2954. [Google Scholar] [CrossRef]
  3. Deng, P.; Mo, W.F.; Ouyang, Z.Q.; Tang, C.L.; Luo, B.H.; Bai, Z.H. Mechanical properties and corrosion behaviors of (Sc, Zr) modified Al-Cu-Mg alloy. Mater. Charact. 2023, 196, 112619. [Google Scholar] [CrossRef]
  4. Zhao, X.Y.; Liu, W.S.; Xiao, D.H.; Ma, Y.Z.; Huang, L.P.; Tang, Y. A critical review: Crystal structure, evolution and interaction mechanism with dislocations of nano precipitates in Al-Li alloys. Mater. Design 2022, 217, 110629. [Google Scholar] [CrossRef]
  5. Ren, Z.X.; Qin, J.Q.; Liu, R.P.; Zhang, X.Y. Vacancy mediated alloying strengthening effects on Al/Al3Zr interface and stabilization of L12-Al3Zr: A first-principles study. J. Alloys Compd. 2020, 825, 153825. [Google Scholar] [CrossRef]
  6. Song, X.; Fu, B.; Chen, X.; Zhang, J.; Liu, T.; Yang, C.; Ye, Y. Effect of internal defects on tensile strength in SLM additively-manufactured aluminum alloys by simulation. Chinese J. Aeronaut. 2023, 36, 485–497. [Google Scholar] [CrossRef]
  7. Zhang, Z.L.; Ghasemi, A.; Koutná, N.; Xu, Z.; Grünstäudl, T.; Song, K.X.; Holec, D.; He, Y.B.; Mayrhofer, P.H.; Bartosik, M. Correlating point defects with mechanical properties in nanocrystalline TiN thin films. Mater. Design 2021, 207, 109844. [Google Scholar] [CrossRef]
  8. Ying, T.; Gu, L.D.; Tang, X.Y.; Wang, J.Y.; Zeng, X.Q. Effect of Sc microalloying on microstructure evolution and mechanical properties of extruded Al-Zn-Mg-Cu alloys. Mater. Sci. Eng. A 2022, 831, 142197. [Google Scholar] [CrossRef]
  9. Wang, Y.X.; Wu, G.H.; Zhang, L.; Guo, Y.J.; Wang, C.L.; Li, L.B.; Xiong, X.M. Enhanced strength and ductility in sand-cast Al-Li-Cu-Mg-Zr alloy via synergistic microalloying with Sc and Ti. J. Alloys Compd. 2023, 962, 170954. [Google Scholar] [CrossRef]
  10. Zhang, J.S.; Wu, G.H.; Zhang, L.; Shi, C.C.; Wang, Y.X.; Sun, B.D.; Li, L.B. Role of Ti in the microstructure evolutions and mechanical properties of cast Al-2.5Li-1.5Cu-1Zn-0.5Mg(−0.2Zr) alloys. J. Alloys Compd. 2022, 899, 163320. [Google Scholar] [CrossRef]
  11. Dong, T.H.; Zhang, X.D.; Yang, L.M.; Wang, F. Investigation of the structural stability, mechanical, and thermodynamic properties of Hf5BSi3 silicide with vacancies defects. Vacuum 2021, 191, 110349. [Google Scholar] [CrossRef]
  12. Jiang, M.; Guo, X.C.; Zu, X.T.; Singh, C.V. The effects of point defects on thermal-mechanical properties of BiCuOTe: A first-principles study. Phys. Chem. Chem. Phys. 2023, 25, 10715–10725. [Google Scholar] [CrossRef]
  13. Fan, T.; Ruan, Z.; Nie, B.; Liao, Y.; Huang, B.; Xu, Z.; Wu, Y.; Hu, T.; Chen, D. First-Principles Investigation of Point Defects on the Thermal Conductivity and Mechanical Properties of Aluminum at Room Temperature. Coatings 2023, 13, 1357. [Google Scholar] [CrossRef]
  14. Muzy, J.; Fivel, M.; Labor, S.; Guignier, D.; Baruchel, J.; Nhi Tran Thi Caliste, T.; Nagirnyi, V.; Romet, I.; Duffar, T. Influence of growth process and crystal defects on sapphire brittleness. J. Cryst. Growth 2023, 618, 127327. [Google Scholar] [CrossRef]
  15. Ouahrani, T.; Boufatah, R.M.; Benaissa, M.; Morales-García, Á.; Badawi, M.; Errandonea, D. Effect of intrinsic point defects on the catalytic and electronic properties of Cu2WS4 single layer:Ab initio calculations. Phys. Rev. Mater. 2023, 7, 025403. [Google Scholar] [CrossRef]
  16. Wang, Y.; Meng, Y.N.; Wang, J.S.; Zhang, C.; Huang, H.B. Mechanical properties of defective L12-Al3X (X = Sc, Lu) phase: A first-principles study. J. Rare Earths 2021, 39, 217–224. [Google Scholar] [CrossRef]
  17. Zhang, M.S.; Wang, J.S.; Wang, B.; Xue, C.P.; Liu, X. Quantifying the effects of Sc and Ag on the microstructure and mechanical properties of Al-Cu alloys. Mater. Sci. Eng. A 2022, 831, 142355. [Google Scholar] [CrossRef]
  18. Wang, L.Y.; Ye, B.; Bai, Y.; Zhao, B.B.; Ding, W.J. Effect of Zr and Sc micro-additions on the microstructure and mechanical properties of as-cast Al-5Ce alloy. Mater. Sci. Eng. A 2021, 822, 141654. [Google Scholar] [CrossRef]
  19. Xue, C.P.; Wang, S.; Zhang, Y.X.; Tian, G.Y.; Yang, X.H.; Chang, X.X.; Ke, Y.B.; Xie, Z.H.; Wang, J.S. Uncovering the kinetics of Li-rich clusters and monodisperse core–shell Al3(Zr, Sc) structures in Al-Li-Cu alloys. Mater. Sci. Eng. A 2023, 881, 145393. [Google Scholar] [CrossRef]
  20. Ekaputra, C.N.; Rakhmonov, J.U.; Weiss, D.; Mogonye, J.E.; Dunand, D.C. Microstructure and mechanical properties of cast Al-Ce-Sc-Zr-(Er) alloys strengthened by Al11Ce3 micro-platelets and L12 Al3(Sc,Zr,Er) nano-precipitates. Acta Mater. 2022, 240, 118354. [Google Scholar] [CrossRef]
  21. Jung, J.G.; Farkoosh, A.R.; Seidman, D.N. Microstructural and mechanical properties of precipitation-strengthened Al-Mg-Zr-Sc-Er-Y-Si alloys. Acta Mater. 2023, 257, 119167. [Google Scholar] [CrossRef]
  22. Zhang, X.J.; Wang, H.W.; Yan, B.; Zou, C.M.; Wei, Z.J. The effect of grain refinement and precipitation strengthening induced by Sc or Er alloying on the mechanical properties of cast Al-Li-Cu-Mg alloys at elevated temperatures. Mater. Sci. Eng. A 2021, 822, 141641. [Google Scholar] [CrossRef]
  23. Yi, M.; Zhang, P.; Yang, C.; Cheng, P.M.; Guo, S.W.; Liu, G.; Sun, J. Improving creep resistance of Al-12 wt.% Ce alloy by microalloying with Sc. Scr. Mater. 2021, 198, 113838. [Google Scholar] [CrossRef]
  24. Wang, J.; Peng, J.; Liu, L.X.; Zhang, F.; Peng, J.H.; Tang, H.; Zheng, J.; An, S.L. Study on the Influences of Adding Rare Earth Ce on the Precipitation Behaviors of TiN Inclusions in 20CrMnTi. Materials 2023, 16, 5598. [Google Scholar] [CrossRef] [PubMed]
  25. Bao, D.H.; Cheng, G.G.; Huang, Y.; Qiao, T. Effect of Rare Earth Ce on the Solidification Structure of Fe–18Cr–0.8Si Ferritic Stainless Steel. Metall. Mater. Trans. B 2023, 54, 944–958. [Google Scholar] [CrossRef]
  26. Fang, H.C.; Chen, K.H.; Chen, X.; Chao, H.; Peng, G.S. Effect of Cr, Yb and Zr additions on localized corrosion of Al-Zn-Mg-Cu alloy. Corros. Sci. 2009, 51, 2872–2877. [Google Scholar] [CrossRef]
  27. Fang, H.C.; Chen, K.H.; Zhang, Z.; Zhu, C.J. Effect of Yb additions on microstructures and properties of 7A60 aluminum alloy. Trans. Nonferr. Metal. Soc. 2008, 18, 28–32. [Google Scholar] [CrossRef]
  28. Peng, G.S.; Chen, K.H.; Fang, H.C.; Chen, S.Y. Effect of Cr and Yb additions on microstructure and properties of low copper Al-Zn-Mg-Cu-Zr alloy. Mater. Des. (1980–2015) 2012, 36, 279–283. [Google Scholar] [CrossRef]
  29. Fang, H.C.; Chen, K.H.; Chen, X.; Huang, L.P.; Peng, G.S.; Huang, B.Y. Effect of Zr, Cr and Pr additions on microstructures and properties of ultra-high strength Al-Zn-Mg-Cu alloys. Mater. Sci. Eng. A 2011, 528, 7606–7615. [Google Scholar] [CrossRef]
  30. Tian, J.Z.; Zhao, Y.H.; Hou, H.; Wang, B. The Effect of Alloying Elements on the Structural Stability, Mechanical Properties, and Debye Temperature of Al3Li: A First-Principles Study. Materials 2018, 11, 1471. [Google Scholar] [CrossRef]
  31. Jia, Q.B.; Rometsch, P.; Kürnsteiner, P.; Chao, Q.; Huang, A.J.; Weyland, M.; Bourgeois, L.; Wu, X.H. Selective laser melting of a high strength Al Mn Sc alloy: Alloy design and strengthening mechanisms. Acta Mater. 2019, 171, 108–118. [Google Scholar] [CrossRef]
  32. Michi, R.A.; Plotkowski, A.; Shyam, A.; Dehoff, R.R.; Babu, S.S. Towards high-temperature applications of aluminium alloys enabled by additive manufacturing. Int. Mater. Rev. 2021, 67, 298–345. [Google Scholar] [CrossRef]
  33. Guo, Y.W.; Wei, W.; Shi, W.; Zhang, B.; Zhou, X.R.; Wen, S.P.; Wu, X.L.; Gao, K.Y.; Rong, L.; Huang, H.; et al. Effect of Er and Zr additions and aging treatment on grain refinement of aluminum alloy fabricated by laser powder bed fusion. J. Alloys Compd. 2022, 912, 165237. [Google Scholar] [CrossRef]
  34. Wang, S.; Wang, J.S.; Zhang, C.; Xue, C.P. Continuous precipitate modes of the δ′-Al3Li phase in Al-Li alloys. J. Alloys Compd. 2022, 904, 163800. [Google Scholar] [CrossRef]
  35. Hohenberg, P.; Kohn, W. Inhomogeneous electron gas. Phys. Rev. 1964, 136, B864. [Google Scholar] [CrossRef]
  36. Kohn, W.; Sham, L.J. Self-consistent equations including exchange and correlation effects. Phys. Rev. 1965, 140, A1133. [Google Scholar] [CrossRef]
  37. Alemany, M.M.G.; Gallego, L.J.; González, D.J. Kohn-Shamab initio molecular dynamics study of liquid Al near melting. Phys. Rev. B 2004, 70, 134206. [Google Scholar] [CrossRef]
  38. Kresse, G.; Furthmüller, J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 1996, 6, 15–50. [Google Scholar] [CrossRef]
  39. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169–11186. [Google Scholar] [CrossRef]
  40. Kresse, G.; Hafner, J. Ab initio molecular dynamics for liquid metals. Phys. Rev. B 1993, 47, 558–561. [Google Scholar] [CrossRef]
  41. Kresse, G.; Furthmüller, J.; Hafner, J. Theory of the crystal structures of selenium and tellurium: The effect of generalized-gradient corrections to the local-density approximation. Phys. Rev. B 1994, 50, 13181–13185. [Google Scholar] [CrossRef]
  42. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 3–62. [Google Scholar] [CrossRef]
  43. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Appl. Phys. Lett. 1996, 77, 3865. [Google Scholar] [CrossRef]
  44. Perdew, J.P.; Wang, Y. Accurate and simple analytic representation of the electron-gas correlation energy. Phys. Rev. B 1992, 45, 13244. [Google Scholar] [CrossRef]
  45. Perdew, J.P.; Chevary, J.A.; Vosko, S.H.; Jackson, K.A.; Pederson, M.R.; Singh, D.J.; Fiolhais, C. Atoms, molecules, solids, and surfaces: Applications of the generalized gradient approximation for exchange and correlation. Phys. Rev. B 1992, 46, 6671. [Google Scholar] [CrossRef]
  46. Ceperley, D.M.; Alder, B.J. Ground state of the electron gas by a stochastic method. Appl. Phys. Lett. 1980, 45, 566. [Google Scholar] [CrossRef]
  47. Perdew, J.P.; McMullen, E.R.; Zunger, A. Density-functional theory of the correlation energy in atoms and ions: A simple analytic model and a challenge. Phys. Rev. A 1981, 23, 2785. [Google Scholar] [CrossRef]
  48. White, J.A.; Bird, D.M. Implementation of gradient-corrected exchange-correlation potentials in Car-Parrinello total-energy calculations. Phys. Rev. B 1994, 50, 4954–4957. [Google Scholar] [CrossRef]
  49. Jiang, C.; Gleeson, B. Site preference of transition metal elements in Ni3Al. Scr. Mater. 2006, 55, 433–436. [Google Scholar] [CrossRef]
  50. Jiang, C.; Sordelet, D.J.; Gleeson, B. Site preference of ternary alloying elements in Ni3Al: A first-principles study. Acta Mater. 2006, 54, 1147–1154. [Google Scholar] [CrossRef]
  51. Lozovoı, A.Y.; Ponomarev, K.V.; Vekilov, Y.K.; Korzhavyı, P.A.; Abrikosov, I.A. First-principles investigation of thermal point defects in B2 NiAl. Phys. Solid State 1999, 41, 1630–1636. [Google Scholar] [CrossRef]
  52. Eriş, R.; Akdeniz, M.V.; Mekhrabov, A.O. The site preferences of transition elements and their synergistic effects on the bonding strengthening and structural Stability of γ′-Ni3Al Precipitates in Ni-Based Superalloys: A First-Principles Investigation. Metall. Mater. Trans. A 2021, 52, 2298–2313. [Google Scholar] [CrossRef]
  53. Pius, K.K.; Ongwen, N.O.; Mageto, M.; Odari, V.; Gaitho, F.M. Mechanical properties of Al-Mg-Si Alloys (6xxx Series): A DFT-Based Study. Alloys 2023, 2, 213–226. [Google Scholar] [CrossRef]
  54. Ali, H.; Chen, R.; Wu, B.; Xie, T.L.; Weng, L.J.; Wen, J.S.; Yao, Q.P.; Su, L.J.; Zhao, Y.; Zhao, P.H.; et al. The site preference and doping effect on mechanical properties of Ni3Al-based γ′ phase in superalloys by combing first-principles calculations and thermodynamic model. Arab. J. Chem. 2022, 15, 104278. [Google Scholar] [CrossRef]
  55. Polychronopoulou, K.; AlKhoori, S.; AlBedwawi, S.; Alareeqi, S.; Hussien, A.G.S.; Vasiliades, M.A.; Efstathiou, A.M.; Petallidou, K.C.; Singh, N.; Anjum, D.H.; et al. Decoupling the Chemical and Mechanical Strain Effect on Steering the CO2 Activation over CeO2-Based Oxides: An Experimental and DFT Approach. ACS Appl. Mater. Inter. 2022, 14, 33094–33119. [Google Scholar] [CrossRef]
  56. Li, Y.L.; Qiang, W.J. Compositional effects on the antiphase boundary energies in B2-type NiTi and NiTi-based high-entropy intermetallics. Mater. Chem. Phys. 2024, 311, 128549. [Google Scholar] [CrossRef]
  57. Singh, M.; Chakraborty, B. 2D BN-biphenylene: Structure stability and properties tenability from a DFT perspective. Phys. Chem. Chem. Phys. 2023, 25, 16018–16029. [Google Scholar] [CrossRef]
  58. Hu, J.Q.; Xie, M.; Zhang, J.M.; Liu, M.M.; Yang, Y.C.; Chen, Y.T. First principles study of Au-Sn intermetallic compounds. Acta Phys. Sin. 2013, 62, 247102. [Google Scholar]
  59. 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]
  60. Connétable, D.; Thomas, O. First-principles study of nickel-silicides ordered phases. J. Alloys Compd. 2011, 509, 2639–2644. [Google Scholar] [CrossRef]
  61. Wang, S.C.; Wang, F.Z. Effect of Mn, N co-doped LiFePO4 on electrochemical and mechanical properties: A DFT study. J. Mol. Graph. Model. 2023, 125, 108604. [Google Scholar] [CrossRef] [PubMed]
  62. Yao, J.G.; Pan, R.K.; Yin, D.F.; Jiang, Y.; Wang, H. Prediction on elastic and thermal properties of defective L12–Al3Li intermetallics. Indian J. Phys. 2018, 92, 1083–1089. [Google Scholar] [CrossRef]
  63. Zhang, X.D.; Jiang, W. Mechanical and thermodynamic properties of Al3Tm and Al3Lu intermetallics under high pressure from first-principles calculations. Chin. Phys. B 2016, 25, 026301. [Google Scholar] [CrossRef]
  64. Wang, V.; Xu, N.; Liu, J.C.; Tang, G.; Geng, W.T. VASPKIT: A User-friendly Interface Facilitating High-throughput Computing and Analysis Using VASP Code. Comput. Phys. Commun. 2021, 267, 08269. [Google Scholar] [CrossRef]
  65. Røyset, J.; Ryum, N. Scandium in aluminium alloys. Int. Mater. Rev. 2013, 50, 19–44. [Google Scholar] [CrossRef]
  66. Oliver, W.C.; Pharr, G.M. Measurement of hardness and elastic modulus by instrumented indentation: Advances in understanding and refinements to methodology. J. Mater. Res. 2011, 19, 3–20. [Google Scholar] [CrossRef]
  67. Hu, W.C.; Liu, Y.; Li, D.J.; Zeng, X.Q.; Xu, C.S. Mechanical and thermodynamic properties of Al3Sc and Al3Li precipitates in Al-Li-Sc alloys from first-principles calculations. Physica B 2013, 427, 85–90. [Google Scholar] [CrossRef]
  68. Jahnátek, M.; Krajčí, M.; Hafner, J. Response of trialuminides to [110] uniaxial loading: An ab initio study forAl3(Sc,Ti,V). Phys. Rev. B 2007, 76, 014110. [Google Scholar] [CrossRef]
  69. Senkov, O.N.; Shagiev, M.R.; Senkova, S.V.; Miracle, D.B. Precipitation of Al3(Sc,Zr) particles in an Al-Zn-Mg-Cu-Sc-Zr alloy during conventional solution heat treatment and its effect on tensile properties. Acta Mater. 2008, 56, 3723–3738. [Google Scholar] [CrossRef]
  70. Jo, H.H.; Fujikawa, S.I. Kinetics of precipitation in Al-Sc alloys and low temperature solid solubility of scandium in aluminium studied by electrical resistivity measurements. Mater. Sci. Eng. A 1993, 171, 151–161. [Google Scholar] [CrossRef]
  71. Murray, J.; Peruzzi, A.; Abriata, J.P. The Al-Zr (aluminum-zirconium) system. J. Phase Equilib. 1992, 13, 277–291. [Google Scholar] [CrossRef]
  72. Hausler, I.; Kamachali, R.D.; Hetaba, W.; Skrotzki, B. Thickening of T1 Precipitates during Aging of a High Purity Al-4Cu-1Li-0.25Mn Alloy. Materials 2018, 12, 30. [Google Scholar] [CrossRef] [PubMed]
  73. Kang, S.J.; Kim, T.H.; Yang, C.W.; Lee, J.I.; Park, E.S.; Noh, T.W.; Kim, M. Atomic structure and growth mechanism of T1 precipitate in Al-Cu-Li-Mg-Ag alloy. Scr. Mater. 2015, 109, 68–71. [Google Scholar] [CrossRef]
  74. Jiang, B.; Cao, F.H.; Wang, H.S.; Yi, D.Q.; Jiang, Y.; Shen, F.H.; Wang, B.; Liu, H.Q. Effect of aging time on the microstructure evolution and mechanical property in an Al-Cu-Li alloy sheet. Mater. Sci. Eng. A 2019, 740–741, 157–164. [Google Scholar] [CrossRef]
  75. Yoshi-Yama, T.; Hasebe, K.; Mannami, M. Al3Li Superlattice inAl-4.5wt.% Li Alloy. J. Phys. Soc. Jpn. 1968, 25, 908. [Google Scholar] [CrossRef]
  76. Mao, Z.; Chen, W.; Seidman, D.N.; Wolverton, C. First-principles study of the nucleation and stability of ordered precipitates in ternary Al-Sc-Li alloys. Acta Mater. 2011, 59, 3012–3023. [Google Scholar] [CrossRef]
  77. Pearson, W.B. A Handbook of Lattice Spacings and Structures of Metals and Alloys: International Series of Monographs on Metal Physics and Physical Metallurgy; Pergamon Press: New York, NY, USA, 1967; Volume 2. [Google Scholar]
  78. Woodward, C.; Asta, M.; Kresse, G.; Hafner, J. Density of constitutional and thermal point defects inL12Al3Sc. Phys. Rev. B 2001, 63, 094103. [Google Scholar] [CrossRef]
  79. Li, Z.Q.; Tse, J.S. Ab initio studies on the vibrational and thermal properties of Al3Li. Phys. Rev. B 2000, 61, 14531–14536. [Google Scholar] [CrossRef]
  80. Tian, T.; Wang, X.F.; Li, W. Ab initio calculations on elastic properties in L12 structure Al3X and X3Al-type (X = transition or main group metal) intermetallic compounds. Solid State Commun. 2013, 156, 69–75. [Google Scholar] [CrossRef]
  81. Hu, H.; Zhao, M.Q.; Wu, X.Z.; Jia, Z.H.; Wang, R.; Li, W.G.; Liu, Q. The structural stability, mechanical properties and stacking fault energy of Al3Zr precipitates in Al-Cu-Zr alloys: HRTEM observations and first-principles calculations. J. Alloys Compd. 2016, 681, 96–108. [Google Scholar] [CrossRef]
  82. Ghosh, G.; Vaynman, S.; Asta, M.; Fine, M.E. Stability and elastic properties of L12-(Al,Cu)3(Ti,Zr) phases: Ab initio calculations and experiments. Intermetallics 2007, 15, 44–54. [Google Scholar] [CrossRef]
  83. Wang, S.; Yang, X.H.; Wang, J.S.; Zhang, C.; Xue, C.P. Identifying the crystal structure of T1 precipitates in Al-Li-Cu alloys by ab initio calculations and HAADF-STEM imaging. J. Mater. Sci. Technol. 2023, 133, 41–57. [Google Scholar] [CrossRef]
  84. Krug, M.E.; Mao, Z.; Seidman, D.N.; Dunand, D.C. Comparison between dislocation dynamics model predictions and experiments in precipitation-strengthened Al-Li-Sc alloys. Acta Mater. 2014, 79, 382–395. [Google Scholar] [CrossRef]
  85. Chen, L.; Wang, Q.; Xiong, L.; Gong, H.R. Mechanical properties and point defects of MC (M = Ti, Zr) from first-principles calculation. J. Alloys Compd. 2018, 747, 972–977. [Google Scholar] [CrossRef]
  86. Sun, S.P.; Li, X.P.; Yang, J.; Wang, H.J.; Jiang, Y.; Yi, D.Q. Point defect concentrations of L12-Al3X (Sc, Zr, Er). Rare Metals 2016, 37, 699–706. [Google Scholar] [CrossRef]
  87. Gao, M.; Deng, Y.H.; Wen, D.D.; Zhao, H.P. A first-principles study on the site occupancy behavior of transition metals in L12-Al3Li phase. Mater. Res. Express 2020, 7, 046519. [Google Scholar] [CrossRef]
  88. Lutsko, J.F.; Wolf, D.; Phillpot, S.R.; Yip, S. Molecular-dynamics study of lattice-defect-nucleated melting in metals using an embedded-atom-method potential, Phys. Rev. B: Condens. Matter Mater. Phys. 1989, 40, 2841–2855. [Google Scholar] [CrossRef]
  89. Shi, Z.M.; Wang, Q.; Zhao, G.; Zhang, R.Y. Effects of erbium modification on the microstructure and mechanical properties of A356 aluminum alloys. Mater. Sci. Eng. A 2015, 626, 102–107. [Google Scholar] [CrossRef]
  90. Spedding, F.H.; Daane, A.H.; Herrmann, K.W. The crystal structures and lattice parameters of high-purity scandium, yttrium and the rare earth metals. Acta Crystallogr. 1956, 9, 559–563. [Google Scholar] [CrossRef]
  91. Brandes, E.A.; Brook, G.B. Smithells Metal Reference Book; Butterworth-Heinemann: Oxford, UK, 1992. [Google Scholar]
  92. Greenwood, N.N.; Earnshaw, A. Chemistry of the Elements; Butterworth-Heinemann: Oxford, UK, 2012. [Google Scholar]
  93. Gu, J.L.; Bai, J.; Zhu, Y.; Qin, Y.; Gu, H.M.; Zhai, Y.C.; Ma, P.H. First-principles study of the influence of doping elements on phase stability, crystal and electronic structure of Al2Cu (θ) phase. Comput. Mater. Sci. 2016, 111, 328–333. [Google Scholar] [CrossRef]
  94. Ahmad, S.; Shafiq, M.; Ahmad, R.; Jalali-Asadabadi, S.; Ahmad, I. Strongly correlated intermetallic rare-earth monoaurides (Ln-Au): Ab-initio study. J. Rare Earth. 2018, 36, 1106–1111. [Google Scholar] [CrossRef]
  95. Zhang, X.; Gui, W.H.; Zeng, Q.F. First-principles study of structural, mechanical, and thermodynamic properties of cubic Y2O3 under high pressure. Ceram. Int. 2017, 43, 3346–3355. [Google Scholar] [CrossRef]
  96. Yang, Y.; Lu, H.; Yu, C.; Chen, J.M. First-principles calculations of mechanical properties of TiC and TiN. J. Alloys Compd. 2009, 485, 542–547. [Google Scholar] [CrossRef]
  97. Gschneidner, K.A. Physical Properties and Interrelationships of Metallic and Semimetallic Elements; Academic Press: Cambridge, MA, USA, 1964; pp. 275–426. [Google Scholar]
  98. Yu, H.L.; Duan, X.H.; Ma, Y.J.; Zeng, M. First principles study of Al-Li intermetallic compounds. Chin. J. Chem. Phys. 2012, 25, 659–665. [Google Scholar] [CrossRef]
  99. Zhang, X.; Jiang, W. First-principles investigations on vibrational, thermodynamic, mechanical properties and thermal conductivity of L12Al3X (X = Sc, Er, Tm, Yb) intermetallics. Phys. Scr. 2015, 90, 065701. [Google Scholar] [CrossRef]
  100. Duan, Y.H.; Sun, Y.; Peng, M.J.; Zhou, S.G. Ab-initio investigations on elastic properties in L12 structure Al3Sc and Al3Y under high pressure. J. Alloys Compd. 2014, 585, 587–593. [Google Scholar] [CrossRef]
  101. Kim, G.; Diao, H.; Lee, C.; Samaei, A.T.; Phan, T.; de Jong, M.; An, K.; Ma, D.; Liaw, P.K.; Chen, W. First-principles and machine learning predictions of elasticity in severely lattice-distorted high-entropy alloys with experimental validation. Acta Mater. 2019, 181, 124–138. [Google Scholar] [CrossRef]
Figure 1. The crystal structures of pure L12 cubic of Al3X. The blue and green balls represent Al and X atoms, respectively.
Figure 1. The crystal structures of pure L12 cubic of Al3X. The blue and green balls represent Al and X atoms, respectively.
Metals 14 00030 g001
Figure 2. The crystal structures of defective off-stoichiometric Al3X supercells with four types of point defects, (a1) Al vacancy (VAl), (a2) Al antisite (XAl), (a3) X vacancy (VX), and (a4) X antisite (AlX). (b) The mean distance of different atomic pairs in Al3X. The blue and green balls represent Al and X atoms, respectively.
Figure 2. The crystal structures of defective off-stoichiometric Al3X supercells with four types of point defects, (a1) Al vacancy (VAl), (a2) Al antisite (XAl), (a3) X vacancy (VX), and (a4) X antisite (AlX). (b) The mean distance of different atomic pairs in Al3X. The blue and green balls represent Al and X atoms, respectively.
Metals 14 00030 g002
Figure 3. The formation enthalpies of stoichiometric Al3X and Al-rich and X-rich off-stoichiometric Al3X alloys as a function of the concentration of the Al atom. The previous data from reference Gao et al. [87], Yao et al. [62], and Wang et al. [16].
Figure 3. The formation enthalpies of stoichiometric Al3X and Al-rich and X-rich off-stoichiometric Al3X alloys as a function of the concentration of the Al atom. The previous data from reference Gao et al. [87], Yao et al. [62], and Wang et al. [16].
Metals 14 00030 g003
Figure 4. The radial distribution functions (RDFs) for the Al-Al, X-X, and Al-X atomic pairs in defective off-stoichiometric Al3X: (1a3a) VAl, (1b3b) XAl, (1c3c) VX, and (1d3d) AlX.
Figure 4. The radial distribution functions (RDFs) for the Al-Al, X-X, and Al-X atomic pairs in defective off-stoichiometric Al3X: (1a3a) VAl, (1b3b) XAl, (1c3c) VX, and (1d3d) AlX.
Metals 14 00030 g004
Figure 5. The four possible position sites of Er-doped in off-stoichiometric Al3X supercells: (a) ErAl, (b) ErX + XAl, (c) ErX, and (d) ErAl + AlX. (Al in blue, X in green, and Er in grey with red circles).
Figure 5. The four possible position sites of Er-doped in off-stoichiometric Al3X supercells: (a) ErAl, (b) ErX + XAl, (c) ErX, and (d) ErAl + AlX. (Al in blue, X in green, and Er in grey with red circles).
Metals 14 00030 g005
Figure 6. The radial distribution functions (RDFs) for the Al-Al, X-X, Al-X, Er-Al, and Er-Li atomic pairs in defective off-stoichiometric Al3X with (1a3a) ErAl, (1b3b) ErX + XAl, (1c3c) ErX, and (1d3d) ErAl + AlX.
Figure 6. The radial distribution functions (RDFs) for the Al-Al, X-X, Al-X, Er-Al, and Er-Li atomic pairs in defective off-stoichiometric Al3X with (1a3a) ErAl, (1b3b) ErX + XAl, (1c3c) ErX, and (1d3d) ErAl + AlX.
Metals 14 00030 g006
Figure 7. Young’s modulus of Al3Sc as a function of indentation depth.
Figure 7. Young’s modulus of Al3Sc as a function of indentation depth.
Metals 14 00030 g007
Figure 8. (a) Bulk modulus B; (b) Shear modulus G; (c) Young’s modulus E; (d) Poisson’s ratios; and (e) Ratios of the bulk modulus to shear modulus (B/G) of pure, defected, and Er-doped L12-Al3X phases.
Figure 8. (a) Bulk modulus B; (b) Shear modulus G; (c) Young’s modulus E; (d) Poisson’s ratios; and (e) Ratios of the bulk modulus to shear modulus (B/G) of pure, defected, and Er-doped L12-Al3X phases.
Metals 14 00030 g008
Figure 9. The Young’s modulus of a 32-atom doped off-stoichiometric Al3Li supercell, the colorless elements indicate that the structure is unstable or no calculation was performed.
Figure 9. The Young’s modulus of a 32-atom doped off-stoichiometric Al3Li supercell, the colorless elements indicate that the structure is unstable or no calculation was performed.
Metals 14 00030 g009
Table 1. Calculated lattice parameters (a0, Å) and bulk modulus (B0, GPa) of Al3X.
Table 1. Calculated lattice parameters (a0, Å) and bulk modulus (B0, GPa) of Al3X.
PhasesLattice Parameter and Bulk ModuluPBELDAPW91Cal. RefExpt. Ref
Al3Lia0 (Å)4.0273.9674.0323.949LDA [79]4.010 [75]
B0 (GPa)64.03270.0762.4062.9GGA [76] 72.3LDA [76]66.0 [76]
Al3Sca0 (Å)4.1034.0344.1054.038LDA [78]4.101 [77]
B0 (GPa)89.28100.2488.0386.6PW91 [80] 87.3PBE [80]91.7 [78]
Al3Zra0 (Å)4.1064.0414.1104.107PBE [81] 4.110PW91 [82] 4.097 [82]-
B0 (GPa)102.16113.45100.46103.45PW91 [80] 106.0PBE [80]-
Table 2. The formation enthalpies of stoichiometric Al3X and defective off-stoichiometric Al3X structures, and the point defect formation enthalpies of off-stoichiometric Al3X structures.
Table 2. The formation enthalpies of stoichiometric Al3X and defective off-stoichiometric Al3X structures, and the point defect formation enthalpies of off-stoichiometric Al3X structures.
DefectH (eV/atom)Hd (eV/Defect)
PerfectVAlXAlVXAlXVAlXAlVXAlX
Al24Li8Cal.−0.100−0.074−0.101−0.074−0.0810.799−0.0400.8090.606
Ref. [87]−0.100−0.092−0.101−0.092−0.0940.791−0.0880.8230.583
Ref. [62]−0.080−0.054−0.073−0.053−0.060----
Al24Sc8Cal.−0.454−0.421−0.435−0.393−0.3741.0060.5941.8922.552
Ref. [16]−0.463−0.428−0.443−0.399−0.3811.4900.5842.4022.575
Al24Zr8Cal.−0.488−0.474−0.481−0.450−0.4180.4300.2151.1692.238
Ref. [86]-----0.4100.3201.0602.070
Table 3. Formation enthalpies ΔH (eV/atom) of doping Er in Al3X.
Table 3. Formation enthalpies ΔH (eV/atom) of doping Er in Al3X.
Phase H (eV/atom)
ErAlErX + XAlErXErAl + AlX
Al24Li8−0.125−0.189−0.171−0.112
Al24Sc8−0.533−0.554−0.585−0.452
Al24Zr8−0.598−0.617−0.639−0.522
Table 4. The calculated bulk modulus B (GPa), shear modulus G (GPa), Young’s modulus E (GPa), Poisson’s ratios v, and ratios of the bulk modulus to shear modulus (B/G) of Al3X in comparison to experiments.
Table 4. The calculated bulk modulus B (GPa), shear modulus G (GPa), Young’s modulus E (GPa), Poisson’s ratios v, and ratios of the bulk modulus to shear modulus (B/G) of Al3X in comparison to experiments.
PhaseSpeciesBGEvB/G
Al24Li8Perfect63.842.7104.70.2261.495
Ref. [30]62.842.3103.60.2251.484
Ref. [98]62.742.7104.40.2221.468
Ref. [76]62.949.0116.60.1911.284
Ref. [79]66.043.0105.90.2321.535
VAl68.238.797.70.2611.760
LiAl67.440.3100.80.2511.673
VLi65.243.9107.60.2251.484
AlLi67.039.298.30.2551.711
ErAl61.438.395.00.2421.605
ErLi + LiAl63.244.5108.10.2151.421
ErLi66.644.5109.10.2271.498
ErAl + AlLi64.037.594.20.2551.707
Al24Sc8Perfect88.269.4164.90.1891.272
Ref. [76]91.871.7170.70.1901.280
Ref. [78]92.378.7183.90.1681.173
Ref. [99]85.972.2169.10.1711.190
Ref. [80]87.371.2168.00.1791.226
Ref. [100]86.074.5173.50.1641.154
Ref. [78]91.768.2163.90.2021.345
VAl89.966.4159.80.2041.354
ScAl87.967.0160.20.1961.312
VSc84.368.2161.10.1821.237
AlSc85.165.6156.60.1931.297
ErAl89.966.4159.80.2041.354
ErSc + ScAl97.872.8174.90.2021.344
ErSc93.271.1170.10.1961.310
ErAl + AlSc85.165.6156.60.1931.297
Al24Zr8Perfect103.065.8162.70.2371.567
Ref. [80]106.065.9163.80.2431.608
Ref. [82]101.462.9156.40.2431.612
Ref. [82]99.682.1193.10.1771.213
Ref. [81]102.666.7164.50.2331.538
VAl110.070.1173.40.2371.570
ZrAl123.565.0165.90.2761.900
VZr100.470.8172.10.2141.417
AlZr100.561.6153.50.2451.630
ErAl99.762.3154.70.2411.600
ErZr + ZrAl100.265.0160.40.2331.541
ErZr100.365.1160.60.2331.541
ErAl + AlZr96.363.1155.40.2311.527
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

Meng, Y.; Wang, S.; Su, H.; Zhang, C.; Wang, B.; Xue, C.; Wang, J. The Effect of Point Defects on Young’s Modulus of the Off-Stoichiometric Al3X (X = Li, Sc, and Zr) Phases: A First-Principles Study. Metals 2024, 14, 30. https://doi.org/10.3390/met14010030

AMA Style

Meng Y, Wang S, Su H, Zhang C, Wang B, Xue C, Wang J. The Effect of Point Defects on Young’s Modulus of the Off-Stoichiometric Al3X (X = Li, Sc, and Zr) Phases: A First-Principles Study. Metals. 2024; 14(1):30. https://doi.org/10.3390/met14010030

Chicago/Turabian Style

Meng, Yanan, Shuo Wang, Hui Su, Chi Zhang, Bing Wang, Chengpeng Xue, and Junsheng Wang. 2024. "The Effect of Point Defects on Young’s Modulus of the Off-Stoichiometric Al3X (X = Li, Sc, and Zr) Phases: A First-Principles Study" Metals 14, no. 1: 30. https://doi.org/10.3390/met14010030

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