Next Article in Journal
Simultaneous Analysis of Cyanotoxins β-N-methylamino-L-alanine (BMAA) and Microcystins-RR, -LR, and -YR Using Liquid Chromatography–Tandem Mass Spectrometry (LC-MS/MS)
Previous Article in Journal
Antioxidant and Hypolipidemic Activities of Cinnamic Acid Derivatives
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

First-Principles Study of Atomic Diffusion by Vacancy Defect of the L12-Al3M (M = Sc, Zr, Er, Y) Phase

1
School of Materials Science and Hydrogen Energy, Foshan University, Foshan 528000, China
2
Research Institute of Automobile Parts Technology, Hunan Institute of Technology, Hengyang 421002, China
3
Shenyuan Honors of College, Beihang University, Beijing 100191, China
4
Shenzhen Rspower Technology Co., Ltd., Shenzhen 518000, China
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(18), 6727; https://doi.org/10.3390/molecules28186727
Submission received: 17 August 2023 / Revised: 4 September 2023 / Accepted: 13 September 2023 / Published: 21 September 2023

Abstract

:
Atomic diffusion by the vacancy defect of L12-Al3M (M = Sc, Zr, Er, Y) was investigated based on a first-principles calculation. The point defect formation energies were firstly evaluated. Then, the migration energy for different diffusion paths was obtained by the climbing-image nudged elastic band (CI-NEB) method. The results showed that Al atomic and M atomic diffusions through nearest-neighbor jump (NNJ) mediated by Al vacancy (VAl) were, respectively, the preferred diffusion paths in Al3M phases under both Al-rich and M-rich conditions. The other mechanisms, such as six-jump cycle (6JC) and next-nearest-neighbor jump (NNNJ), were energetically inhibited. The order of activation barriers for NNJ(Al-VAl) was Al3Zr < Al3Y < Al3Er < Al3Sc. The Al3Sc phase had high stability with a high self-diffusion activation barrier, while the Al3Zr and Al3Y phases were relatively unstable with a low self-diffusion activation energy. Moreover, the atomic-diffusion behavior between the core and shell layers of L12-Al3M was also further investigated. Zr atoms were prone to diffusion into the Al3Y core layer, resulting in no stable core-shelled Al3(Y,Zr), which well agreed with experimental observation.

1. Introduction

Aluminum alloys containing transition metals (TM) and rare-earth (RE) elements are widely used in aerospace, high-speed trains, and automobiles [1,2,3]. The formation of L12-Al3(RE,TM) nanophases with high thermal stability, such as the core-shelled Al3(Sc,Zr) nanophase, can effectively inhibit the recrystallization process of aluminum alloys, achieving high strength toughness and corrosion resistance performance [2]. Based on the high thermostability and coarsening resistance of L12-Al3(RE,TM) nanophases, Seidman et al. [4,5,6] developed an Al-RE-TM series of high-temperature aluminum alloys.
Due to the high cost of Sc elements, low-cost elements that can form thermally stable Al3M phases were explored to replace Sc elements. A first-principles calculation showed that Er and Yb elements were candidate element to replace Sc elements [7,8]. The Zr element can partially replace Sc element and was considered an ideal element to form the Al3Zr shell structure owing to the Al3Zr/Al interface with low interface energy and coherent strain energy [9]. The typical core-shelled nanophases, such as Al3(Er,Zr) [10] and Al3(Yb,Zr) [11], were introduced in aluminum alloys to replace Al3(Sc,Zr) nanophases. The formation of these core-shelled Al3M phases was explained by the difference in diffusion rates between elements, that is, the rapid diffusion elements were enriched to form a core layer and the slow diffusion elements were segregated to form a shell layer [12]. The core-double-shelled structure that was observed in Al-Sc-Er-Zr alloys contained an Er-enriched core surrounded by a Sc-enriched core and a Zr-enriched outer shell, obtaining a high coarsening resistant and high strength [5,13]. The formation of the core-double-shelled structure well agreed with the prediction by the atomic diffusivity ordering of DEr > DSc > DZr,
The Y element, as a Sc homologous element with similar physical and chemical properties, was a probable candidate element to replace the Sc element [14]. Zhang et al. [15,16] showed that core-shelled Al3(Y,Zr) with a Y-rich core and Zr-rich shell can be formed during the early stage of aging in Al-Y-Zr alloys. The Al3Y phase acted as the heterogeneous core to accelerate the precipitation of Al3Zr, which well agreed with the atomic-diffusion control mechanism. However, atom probe tomography (APT) showed that Y and Zr atoms were randomly distributed in the L12-Al3(Y,Zr), and hybrid structure, rather than core-shelled structure, was observed in the Al3(Y,Zr) phase after long-term aging. The strong binding energy between Y and Zr atoms was assumed to explain the formation of hybrid structure Al3(Y,Zr) [16]. According to the authors’ previous investigation, the core-shelled Al3Y/Al3Zr was thermodynamically unstable due to its high coherent strain energy of Al3Y/Al3Zr. A similar transition from a core-shelled structure to a hybrid structure during long-term aging was also observed in the Al-Yb-Sc alloy [17]. Seidman et al. [17] suggested that the inter-diffusion of Yb and Sc resulted in a uniform distribution of elements throughout the precipitates. Thus, the mechanism of the L12-Al3M phase with core-shelled structure or hybrid structure needed to be answered.
Atomic diffusion, especially mediated by vacancy, was very beneficial for understanding the microstructural stability of the L12-Al3M phase [18]. Although experimental methods were difficult to investigate atomic diffusion in intermetallic compounds [19,20], first-principles calculations can provide new insights into the microscopic mechanisms of atomic diffusion [18,20]. First-principles calculations by Fan [21,22] showed that with the increase of the atomic number, the diffusion rate of rare-earth elements increased from Sc to Y, La, and then decreased to Lu. Shi et al. [23] investigated the atomic diffusion of pure and transition-element (TM)-doped L12-Al3Sc based on first principles and found that under a strong Al-rich condition, the VSc defect obtained low formation energy and the NNJ mechanism mediated by VAl was most favorable for Sc atomic diffusion. TM dopants increased diffusion activation energy for dominant Al3Sc diffusion. However, the atomic diffusion between the core layer and the shell layer for the core-shelled L12-Al3M phase was far from fully understood.
The atomic-diffusion mechanism in the L12-Al3M (M = Sc, Er, Y, Zr) phases was investigated based on the first-principles in the present work. The formation energy of point defects was calculated. Then, migration energy along possible diffusion was analyzed by the climbing-image nudged elastic band (CI-NEB) method [24]. Thus, the diffusion activation energy was obtained. Furthermore, the atomic-diffusion behavior between the core and shell layers for the core-shelled L12-Al3M phase was also further illustrated. The purpose of this study was to reveal the microscopic mechanism of atomic diffusion in the L12-Al3M phases and core-shelled L12-Al3M phase, providing a theoretical guidance for the development of high-performance aluminum alloys containing TM and RE elements.

2. Results and Discussion

2.1. Defect Formation Energy

In order to evaluate the difficulty of point defect formation, the formation energy of point defects in Al3M was calculated as [25]
E f d e f = E t o t a l d e f E t o t a l b u l k i Δ n i μ i
Here, E t o t a l d e f and E t o t a l b u l k are the total energy of the defective supercell and the total energy of the defect-free supercell, respectively. ni represents the number of i atoms (i = Al or M) that increased ( Δ n i > 0 ) or decreased ( Δ n i < 0 ) when defects were formed, and μi is the relative chemical potential of i atoms.
To maintain a stably balanced L12 phase, its chemical potential should meet the following requirement:
3 Δ μ A l + Δ μ M = Δ H f A l 3 M
Here, Δ H f A l 3 M is the formation enthalpy of the unit chemical formula Al3M in the solid state; Δ μ A l and Δ μ M are the differences between the relative chemical potentials of Al and M atoms, respectively, and the chemical potentials of solid simple substance, which can be expressed as
Δ μ A l = μ A l μ A l b u l k Δ μ M = μ M μ M b u l k
Here, μ A l b u l k and μ M b u l k are the chemical potentials of the metals Al and M, respectively, that is, the single-atomic energies in the elemental state.
In order to avoid the precipitation of the solid elements Al and M, the chemical potential of each atom in the defect phase should be less than that of the solid elements, that is,
Δ μ A l 0 Δ μ M 0
From the phase diagram of each Al-M system, Al3M was in equilibrium with the adjacent pure Al phase, as the non-stoichiometric ratio was Al-rich due to point defects. When the non-stoichiometric ratio was M-rich due to point defects, the Al3M phase was in equilibrium with the adjacent stoichiometric Al2M phase. In order to avoid the formation of the pure Al and Al2M secondary phase, the chemical potential should be limited by the following:
Δ μ A l 0 2 Δ μ A l + Δ μ M Δ H f A l 2 M
Here, Δ H f A l 2 M is the formation enthalpy of the unit chemical formula Al2M in the solid state.
The formation energies of four kinds of defects in the L12-Al3M phase were calculated, as shown in Figure 1. Under the Al-rich condition, the point defects of VSc obtained low formation energy and were the main point defects for the Al3Sc phase, and the point defects of AlEr and VEr were the main point defects for the Al3Er phase. The ScAl and ErAl defects were the main point defects for the Al3Sc phase and Al3Er phase owing to the lowest formation energy under Sc-rich and Er-rich conditions. The point defects of the Al3Sc phase under Al-rich and Sc-rich conditions were well consistent with Ref. [23]. On the other hand, the change of the stoichiometric ratio had little effect on the formation energy of the point defects in the Al3Zr phase and Al3Y phase. The VZr and AlY defects obtained the lowest formation energy for the Al3Zr phase and Al3Y phase regardless of the Al-rich, Zr-rich, and Y-rich conditions. Furthermore, the defect energy of VZr was a negative value, indicating that the Al3Zr phase was inclined to form stable VZr vacancy defects under the Al-rich condition. The formation energy of the YAl antisite was always the highest; thus, it was difficult to form YAl antisite defects in the Al3Y phase. Shi et al. [23] suggested that the point defect formation energy was dependent on the electronic structure and the value of the electronic density of state (DOS) at the Fermi level. However, VAl was the primary point defect near the stoichiometry [23]. Similar vacancy defects were reported in the Ni3Al phase [18], where vacancies defect on the Ni sublattice was the main point defect in the Ni3Al phase.

2.2. Vacancy-Mediated Atomic Migration

2.2.1. Al Atomic Migration

Figure 2a–c show the energy profiles for Al atomic diffusion mediated by VAl in the Al3M phase. The nearest-neighbor site Al around VAl can migrate to VAl through the symmetrical NNJ pathways (denoted as NNJ(Al-VAl)), and the energy profile was symmetrical due to the restoration of the local disordered structures (Figure 2a). The highest energies of the energy profile corresponding to the migration barrier were 0.913 eV, 0.914 eV, 0.647 eV, and 0.622 eV for Al3Sc, Al3Zr, Al3Er, and Al3Y, respectively, indicating that NNJ(Al-VAl) with the low migration barrier was the preferred diffusion path for Al3M phases owing to the direct jump to VAl. Furthermore, the migration barrier of Al3Sc was almost the same as that of Al3Zr, while Al3Er and Al3Y obtained low migration barriers. The different migration barrier for the NNJ path can be attributed to the different M atomic sizes. The Er and Y atoms had large atomic radii; thus, the Al3Er and Al3Y obtained high lattice gaps, where Al atoms can migrate through the large atomic gaps to VAl, obtaining a lower migration barrier.
Al atoms that occupied the next-nearest-neighbor sites of VAl can migrate to VAl through two types of diffusion paths, denoted as NNNJ1 and NNNJ2, as shown in Figure 2b,c. As for the NNNJ1 path, the order of migration energy was Al3Zr > Al3Sc > Al3Er > Al3Y. Al atoms migrated through the quadrangle composed of the nearest-neighbor Al atoms, where the quadrangle gap became large with the increase in M atomic radius; correspondingly, the migration energy decreased with the increase in M atomic radius. However, the migration energies of the NNNJ2 path were almost the same and significantly increased owing to the high density of the quadrangle with two Al atoms and two M atoms. Compared with the NNJ migration, the Al atoms through the NNNJ1 and NNNJ2 paths needed to cross the quadrangle composed of four neighboring atoms, illustrating higher migration energy [23]. Therefore, the tendency of Al migration by NNNJ was very low.
The diffusion of Al atoms was also mediated by VM, including the NNJ and ASB migration paths. Figure 2d shows the migration-energy profile for the NNJ path of Al atom mediated by VM (denoted as NNJ(Al-VM)), where the migration barriers mediated by VM were still higher than that by VAl. Meanwhile, the final state of migration was unstable due to the local disorder by the migration of Al atoms to VM. In this sense, NNJ(Al-VM) was not the preferred migration path. As for the ASB migration (Figure 2e), AlM occupied the next-nearest-neighbor site of VM, and the nearest-neighbor Al atoms migrated to VM, forming VAl (step 1). Then, AlM migrated to the newly formed VAl (step 2), denoted as ASB(AlM-Al-VM). The order of migration energy for AS and ASB was similar to that for NNNJ1 as Al3Zr > Al3Sc > Al3Er > Al3Y, which can be explained by the different M atomic radii. Although the disorder of the migration final state was restored to the migration initial state, the ASB migration of Al atoms mediated by VM was restricted due to the higher migration barrier of the NNJ migration by VAl. From the above discussion, the NNJ(Al-VAl) path had the lowest migration energy and was the main migration pathway for Al atomic migration.

2.2.2. M Atomic Migration

The migration of M atoms mediated by VAl included the NNJ, AS, ASB, and 6JC paths. The migration of the nearest-neighbor M atoms jumped to VAl, denoted as NNJ(M-VAl), as shown in Figure 3a. The migration barriers of NNJ(M-VAl) were far higher than that of NNJ(Al-VAl). Meanwhile, the final state of migration was unstable due to the local disorder with the formation of MAl and VM defects. However, the Y atoms needed to overcome the increasing energy barrier during the migration process, suggesting that this migration path of Y atoms was energetically prohibited.
The AS migration path was another M diffusion path, where the MAl atoms directly migrated to VAl (denoted as AS(MAl-VAl)), as shown in Figure 3b. The migration barriers of the AS(MAl-VAl) path were very low compared with that of NNJ(M-VAl). Except for the Al3Sc phase, the final state energies of the Al3Zr, Al3Er, and Al3Y phases were negative, which was inconsistent with the fact that the local disorder restored to their initial state after M atomic migration. It suggested that the AS(MAl-VAl) path did not exist for the Al3Zr, Al3Er, and Al3Y phases in terms of energy. Furthermore, the AS(ScAl-VAl) migration path was limited to some extent due to the difficulty of coexisting VAl and ScAl in nearby locations.
For the ASB migration path, MAl atoms and VAl occupied the site of Al atoms (Figure 3c). The nearest-neighbor M firstly jumped to VAl, newly forming a VM and MAl atom; then, the MAl atom migrated to the new VM vacancy, resulting in a VAl (denoted as ASB(MAl-M-VAl)). Obviously, the local disorder of the final state was consistent with that of the initial state; therefore, the migration-energy curve of the M atoms was symmetric. These two migration steps corresponded to the two saddle-curve characteristics of M atomic-migration energy. The Zr atomic migration by the ASB path in the Al3Zr phase obtained the maximum migration energy with 1.754 eV, which was lower than that of NNJ(M-VAl) with 2.059 eV. Meanwhile, the migration barriers of the Er, Y, and Sc atoms were nearly the same for the ASB path. However, the ASB pathway was limited due to the simultaneous presence of both MAl and VAl defects.
The 6JC path of M atomic migration mediated by VAl consisted of the straight 6JC path and the bent 6JC path, as shown in Figure 3d,e. The six steps in the straight 6JC and bent 6JC were described in detail in Ref. [23]. The energy curves of M diffusion by straight 6JC and bent 6JC were symmetric due to the restoration of local disorder after the six-step migration process. The first step (M-VAl) in the straight 6JC and bent 6JC paths was similar to that of NNJ(M-VAl). In the first step, the migration energy of the Y atom increased during the migration process, while the migration energy of the Sc and Zr elements decreased after the migration to VAl. Thus, the Y atomic diffusion obtained far higher migration energy than the Er, Sc, and Zr atoms. The high migration barriers for the Y, Er, Sc, and Zr atoms indicated the straight 6JC and bent 6JC paths were energetically restricted.
The migration of M atoms mediated by VM is also shown in Figure 3f. The M atom at the next-nearest-neighbor site jumped to VM (denoted as NNNJ(M-VM)), and the energy profile of the M diffusion was symmetric owing to the restoration of local disorder. The Zr atom obtained the higher migration barrier than that of the Er, Sc, and Y atoms due to the dense structure of the Al3Zr phase. However, the Sc atomic diffusion had the lowest migration barrier, which did not agree with the effect of atomic size. Therefore, the migration barrier was related not only to the atomic radius of M but also to the electronic structure of the M atom [23]. Furthermore, the migration barriers of M atoms for NNNJ(M-VM) were much higher than that for NNJ(M-VAl) and ASB(MAl-M-VAl), suggesting that the NNNJ(M-VM) migration path was not a preferred migration path.
From the above discussion, except the AS(ScAl-VAl) path being the preferred path for Sc atomic migration, the NNJ(M-VAl) and ASB(MAl-M-VAl) paths contributed to M atomic migration, while the straight 6JC, bent 6JC, and NNNJ(M-VM) paths were energetically prohibited.

2.3. Diffusion Activation Barrier

In the process of vacancy-mediated atomic diffusion, the activation barrier had a decisive influence on atomic diffusion and can be expressed as [23]:
Δ E a = Δ E f d e f e c t + Δ E m
Here, Δ E f d e f e c t is the defect formation energy of the initial state, and Δ E m is the migration barrier for different migration paths.
Table 1 shows that the calculated diffusion activation energies of the Al and M atoms in the Al3M phases. The calculation of the Al3Sc phase generally agreed with that of Ref. [23]. Under the Al-rich condition, the Al atomic-diffusion activation barriers of NNJ(Al-VM) were lower than that of NNJ(Al-VAl) in the Al3M phases, which was attributed to the low VM formation energies under the Al-rich condition. However, the diffusion path of NNJ(Al-VM) was restricted due to the unstable final state (Figure 2d). Thus, NNJ(Al-VAl) diffusion was the main diffusion path for Al atoms under both Al-rich and M-rich conditions.
For M atomic diffusion, NNJ(M-VAl) diffusion obtained the low activation barriers for the Sc, Zr, Er, and Y atoms under both Al-rich and M-rich conditions, thus becoming the energetically preferred diffusion path. The diffusion activation barriers of Y atoms were far higher than that of the Sc, Zr, and Er elements due to their large atomic radii. Although the diffusion barriers of AS(ScAl-VAl) and ASB(ErAl-Er-VAl) were much lower under the M-rich condition, their contribution to Sc atomic and Er atomic diffusions was limited due to the difficulty of coexisting VAl and ScAl, ErAl defects in nearby locations.
The activation barriers of Al atomic and M atomic diffusions mediated by VAl under the M-rich condition were generally lower than that under the Al-rich condition owing to the lower formation energy of VAl. The diffusion activation barriers of Al atoms in the Al3M phase were lower than that of M atoms. Moreover, ASB, NNNJ, straight 6JC, and bent 6JC mechanisms with high diffusion activation barriers were not factually executed for Al atomic and M atomic diffusions under both Al-rich and M-rich conditions. Therefore, the NNJ(Al-VAl) diffusion under the M-rich condition was the most preferred diffusion mechanism in the Al3M phase, and the order of activation barriers was Al3Zr < Al3Y < Al3Er < Al3Sc. It should be noted that these calculations were Al-atomic-self-diffusion activation barriers in the Al3M phase. It implied that the Al3Sc phase had high stability with a high self-diffusion activation energy, while the Al3Zr phase was relatively unstable with a low self-diffusion activation barrier, which well agreed with the fact that the Al3Zr phase transformed from the L12 structure to D023 structure at high temperatures [26].

2.4. Atomic Diffusion of Core-Shelled L12-Al3(N,Zr)

The addition of the Er, Y, Sc, and Zr elements in aluminum alloys typically formed L12-Al3M phases with a core-shelled structure. Due to the low interface energy between the Al3Zr phase and aluminum matrix, the Al3Zr phase tended to form a shell layer, and the Al3N (N = Sc, Er, Y) phase was inclined to form a core layer, where the core-shell structure was denoted as Al3(N,Zr). Atomic diffusion between the Al3N core and Al3Zr shell was investigated in this section. The Zr atomic diffusion to the Al3N core and the N atomic diffusion to the Al3Zr shell were respectively calculated. The previous investigation showed that Zr atoms in the Al3Zr shell reciprocally substituted the site of N atoms in the Al3N core [27].
As described in Section 2.3, NNJ(M-VAl) was the preferred diffusion path for Zr atoms in the Al3N core, where Zr atoms occupied the site of N atoms at the nearest-neighbor site of VAl. The formation energy of VAl was affected by Zr atomic substitution and can be expressed as
E f d e f ( Z r N ) = E t o t a l d e f ( Z r N ) E t o t a l b u l k i Δ n i μ i μ Z r
Here, E t o t a l d e f ( Z r N ) is the total energy of defective supercell with Zr substitution. μZr is the chemical potential of Zr atoms.
For the N atomic substitution for Zr atoms in the Al3Zr shell, the formation energies of VAl defect with N atomic substitution are expressed as
E f d e f ( N Z r ) = E t o t a l d e f ( N Z r ) E t o t a l b u l k i Δ n i μ i μ N
Here, E t o t a l d e f ( N Z r ) is the total energy of the defective supercell with N substitution. μN is the chemical potential of N atoms.
Table 2 shows that the formation energy of the VAl defect with Zr substitution was within that of the pure Al3N phase between Al-rich and N-rich, and Zr substitution had little influence on the formation energy of the VAl defect due to the atomic radius of Al being close to that of Zr. As shown in Table 3, the Sc substitution for Zr atoms slightly increased the formation energy of the VAl defect, while the Er and Y substitutions significantly increased the formation energy of the VAl defect.
Zr atoms were considered to migrate through the NNJ path mediated by VAl in the Al3N core, as shown in Figure 4a. The migration barriers of Zr diffusion in the Al3Sc, Al3Er, and Al3Y cores were 1.658 eV, 1.653 eV, and 1.570 eV, respectively, which were far lower than that in the Al3Zr phase with 2.059 eV. The order of migration barriers for Zr atoms in the Al3N core was Al3Sc > Al3Er >Al3Y. With the increase in the N atomic radius for the Al3N core, Zr atoms easily migrated due to the increase in the lattice gap. Thus, Zr atomic diffusion in the Al3Y core obtained the lowest migration energy. Figure 4b illustrates that the diffusion of Sc, Er, and Y atoms in the Al3Zr shell obtained migration barriers with 2.14 eV, 2.356 eV, and 2.53 eV, respectively, which were higher than that of Zr migration in the Al3Zr shell with 2.059 eV. The atomic-migration energy in the Al3Zr shell was sequentially Y > Er > Sc, which can be attributed to the high diffusion resistance for the large atomic radius.
As the diffusion activation barrier included vacancy formation energy and atomic-migration energy, the diffusion activation energies of Zr atoms in the Al3Sc, Al3Er, and Al3Y cores were 2.852 eV, 2.84 eV, and 2.633 eV, respectively. Additionally, the diffusion activation barriers of the Sc, Er, and Y atoms in the Al3Zr shell were 3.024 eV, 3.492 eV, and 3.889 eV, respectively. Compared with the diffusions of the Sc, Er, and Y atoms into the Al3Zr shell, Zr atoms were more inclined to enter the Al3Sc, Al3Er, and Al3Y cores based on diffusion activation energy. Furthermore, Zr atoms preferred to diffuse into the Al3Y core, while the diffusion of Zr atoms into the Al3Er and Al3Sc cores required higher activation barriers. It revealed that Zr atoms in the Al3Zr shell were inclined to diffuse into the Al3Y core during the subsequent aging process, thus resulting in no typical core-shelled structure. However, Zr atoms were difficult to diffuse into Al3Sc and Al3Er cores due to their high diffusion activation barrier, thus maintaining a typical core-shelled structure.

3. Computational Methods

Based on density functional theory (DFT) [28], first-principles calculations were carried out by Vienna ab initio simulation package (VASP) software [29]. The projector augmented wave (PAW) with the Perdew–Burke–Ernzerh (PBE) method of generalized gradient approximation (GGA) was used to describe the exchange-correlation energy functional between electrons [30]. The electron configuration was described by the Al-3s23p1, Sc-3s23p64s13d2, Zr-4s24p65s14d3, Er-6s25p65d1, and Y-4s24p65s14d2 valence states, respectively. The kinetic energy cutoff of the plane-wave basis and the size of the k-mesh for the Brillouin zone were tested for self-consistent convergence. The geometric structure was optimized by the Monkhorst-Pack k-point grids with linear k-mesh analytical values of less than 0.032π/Å. The total energy was calculated using the linear tetrahedron method with the Blöchl correction when the total energy converged to 10−4 eV/atom. The lattice constants (a0) were predicted as 4.042 Å, 4.103 Å, 4.108 Å, and 4.232 Å for fcc-Al, L12-Al3Sc, L12-Al3Zr, and L12-Al3Er, respectively, which were well consistent with Ref. [31].
There were two sublattices in the L12-Al3M (M = Sc, Zr, Er, Y) unit cell, the Al sublattice located at the 3c position (0,0.5,0.5) and the M sublattice located at the 1a (position (0,0,0). Therefore, there were four types of primary point defects in Al3M, including Al vacancy (VAl), M vacancy (VM), Al antisite (AlM), and M antisite (MAl). In order to reduce vacancy density and limit the interaction between defects, 108 atoms in a 3 × 3 × 3 supercell were used in this calculation.
The diffusion mechanism of the L12-Al3M phase mediated by vacancy included from the nearest-neighbor to complex hopping sequences, such as nearest-neighbor jump (NNJ), next-nearest-neighbor jump (NNNJ), antistructural sublattice (AS), antistructural bridge (ASB), and 6-jump cycle (6JC) [23]. The CI-NEB method [24] was used to calculate the energy profile. A series of atomic positions were inserted between the initial and final states to construct the model, and then, each insertion point model was relaxed until the force threshold at the insertion point was 10−2 eV/Å. By this method, the vacancy diffusion behavior of the L12-Al3M phase was comparatively studied, and the atomic-diffusion interaction between the core layer and the shell layer for the core-shelled L12-Al3M phase was also discussed.

4. Conclusions

Atomic-diffusion mechanisms in L12-Al3M (M = Sc, Er, Y, Zr) phases were investigated based on a first-principles calculation. The main conclusions are summarized as follows:
(1)
NNJ(Al-VAl) and NNJ(M-VAl) diffusions were the energetically preferred diffusion paths under both Al-rich and M-rich conditions. The straight 6JC, bent 6JC, and NNNJ were significantly inhibited owing to their high activation barriers. Other diffusion paths, such as NNJ(Al-VM), AS(ScAl-VAl), and ASB(ErAl-Er-VAl), were limited due to the unstable final-state structure and the difficulty of coexisting VAl and MAl defects.
(2)
The order of activation barriers for NNJ(Al-VAl) was Al3Zr < Al3Y < Al3Er < Al3Sc. The Al3Sc phase had high stability with a high self-diffusion activation barrier, while the Al3Zr phase was relatively unstable with a low self-diffusion activation energy.
(3)
Compared with the diffusion of the Sc, Er, and Y atoms in the Al3Zr shell, Zr atoms were more inclined to diffuse into the Al3Y, Al3Er, and Al3Sc cores, and the activation barriers were as follows: Al3Y < Al3Er < Al3Sc. Thus, Zr atoms were prone to diffuse into the Al3Y core, resulting in no core-shelled structure.

Author Contributions

B.N. and Y.S. conceived and designed the research; S.L., B.L. and J.Z. performed the first-principles calculation; T.F. and D.C. analyzed the experimental data; S.L. wrote the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the R & D plan for key areas in Guangdong Province (2020B010186001), Science and Technology Program of the Ministry of Science and Technology (G2022030060L), Science and technology project in Guangdong (2020b15120093, 2020B121202002), Overseas famous teacher project of Guangdong(BGK46303), Science and technology research project of Foshan (2220001005305), and R and D plan for key areas in Jiangxi Province (20201BBE51009, 20212BBE51012).

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not applicable.

References

  1. Dursun, T.; Soutis, C. Recent developments in advanced aircraft aluminium alloys. Mater. Des. 2014, 56, 862–871. [Google Scholar] [CrossRef]
  2. Ye, J.; Pan, Q.; Liu, B.; Hu, Q.; Qu, L.; Wang, W.; Wang, X. 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.; Ouyang, Z.; Tang, C.; Luo, B.; Bai, Z. Mechanical properties and corrosion behaviors of (Sc, Zr) modified Al-Cu-Mg alloy. Mater. Charact. 2022, 196, 112619. [Google Scholar] [CrossRef]
  4. De Luca, A.; Seidman, D.N.; Dunand, D.C. Effects of Mo and Mn microadditions on strengthening and over-homogenization resistance of nanoprecipitation-strengthened Al-Zr-Sc-Er-Si alloys. Acta Mater. 2019, 165, 1–14. [Google Scholar] [CrossRef]
  5. Booth-Morrison, C.; Dunand, D.C.; Seidman, D.N. Coarsening resistance at 400 °C of precipitation-strengthened Al-Zr-Sc-Er alloys. Acta Mater. 2011, 59, 7029–7042. [Google Scholar] [CrossRef]
  6. Farkoosh, A.R.; Dunand, D.C.; Seidman, D.N. Effects of W and Si microadditions on microstructure and the strength of dilute precipitation-strengthened Al-Zr-Er alloys. Mater. Sci. Eng. A 2020, 798, 140159. [Google Scholar] [CrossRef]
  7. Liu, S.; Liu, F.; Yan, Z.; Nie, B.; Fan, T.; Chen, D.; Song, Y. Nucleation of L12-Al3M (M = Sc, Er, Y, Zr) Nanophases in Aluminum Alloys: A First-Principles Thermodynamics Study. Crystals 2023, 13, 1228. [Google Scholar] [CrossRef]
  8. Lan, J.; Chen, Z.; Liu, L.; Zhang, Q.; He, M.; Li, J.; Peng, X.; Fan, T. The Thermal Properties of L12 Phases in Aluminum Enhanced by Alloying Elements. Metals 2021, 11, 1420. [Google Scholar] [CrossRef]
  9. Song, Y.; Zhan, S.; Nie, B.; Liu, S.; Qi, H.; Liu, F.; Fan, T.; Chen, D. First-principle investigation of the interface properties of the core-shelled L12-Al3M (M = Sc, Zr, Er, Y) Phase. Crystals 2023, 13, 420. [Google Scholar] [CrossRef]
  10. Xue, D.; Wei, W.; Wen, S.; Wu, X.; Shi, W.; Zhou, X.; Gao, K.; Huang, H.; Nie, Z. Microstructural evolution of Al-Mg-Er-Zr alloy by equal channel angular extrusion at room temperature. Mater. Lett. 2023, 334, 133759. [Google Scholar] [CrossRef]
  11. Peng, G.; Chen, K.; Fang, H.; Chen, S. A study of nanoscale Al3 (Zr, Yb) dispersoids structure and thermal stability in Al-Zr-Yb alloy. Mater. Sci. Eng. A 2012, 535, 311–315. [Google Scholar] [CrossRef]
  12. Clouet, E.; Laé, L.; Épicier, T.; Lefebvre, W.; Nastar, M.; Deschamps, A. Complex precipitation pathways in multicomponent alloys. Nat. Mater. 2006, 5, 482–488. [Google Scholar] [CrossRef] [PubMed]
  13. Chen, S.; Li, C.; Lian, G.; Guo, C.; Du, Z. Effect of elastic strain energy on the core-shell structures of the precipitates in Al-Sc-Er alloys. J. Rare Earths 2012, 30, 1276–1280. [Google Scholar] [CrossRef]
  14. Zhang, X.D.; Wang, S.Q. First-principles study of thermodynamic properties and solubility of aluminum-rare-earth intermetallics. Comput. Mater. Sci. 2014, 90, 56. [Google Scholar] [CrossRef]
  15. Zhang, Y.; Gu, J.; Tian, Y.; Gao, H.; Wang, J.; Sun, B. Microstructural evolution and mechanical property of Al-Zr and Al-Zr-Y alloys. Mater. Sci. Eng. A 2014, 616, 132–140. [Google Scholar] [CrossRef]
  16. Gao, H.; Feng, W.; Wang, Y.; Gu, J.; Zhang, Y.; Wang, J.; Sun, B. Structural and compositional evolution of Al3(Zr, Y) precipitates in Al-Zr-Y alloy. Mater. Charact. 2016, 121, 195–198. [Google Scholar] [CrossRef]
  17. Van Dalen, M.E.; Dunand, D.C.; Seidman, D.N. Microstructural evolution and creep properties of precipitation-strengthened Al-0.06 Sc-0.02 Gd and Al-0.06 Sc-0.02 Yb (at.%) alloys. Acta Mater. 2011, 59, 5224–5237. [Google Scholar] [CrossRef]
  18. Zhang, X.; Wang, C.Y. First-principles study of vacancy formation and migration in clean and Re-doped γ′-Ni3Al. Acta Mater. 2009, 57, 224–231. [Google Scholar] [CrossRef]
  19. Yan, K.; Chen, Z.W.; Zhao, Y.N.; Ren, C.C.; Aldeen, A.W. Morphological characteristics of Al3Sc particles and crystallographic orientation relationships of Al3Sc/Al interface in cast Al-Sc alloy. J. Alloys Compd. 2020, 861, 158491. [Google Scholar] [CrossRef]
  20. Li, S.-S.; Li, L.; Han, J.; Wang, C.T.; Xiao, Y.Q.; Jian, X.D.; Qian, P.; Su, Y.J. First-Principles study on the nucleation of precipitates in ternary Al alloys doped with Sc, Li, Zr, and Ti elements. Appl. Surf. Sci. 2020, 526, 146455. [Google Scholar] [CrossRef]
  21. Hu, T.; Ruan, Z.; Fan, T.; Wang, K.; He, K.; Wu, Y. First-principles investigation of the diffusion of TM and the nucleation and growth of L12-Al3TM particles in Al alloys. Crystals 2023, 13, 1032. [Google Scholar] [CrossRef]
  22. Fan, T.; Ruan, Z.; Zhong, F.; Xie, C.; Li, X.; Chen, D.; Tang, P.; Wu, Y. Nucleation and growth of L12-Al3RE particles in aluminum alloys: A first-principles study. J. Rare Earths 2023, 41, 1116–1126. [Google Scholar] [CrossRef]
  23. Shi, T.T.; Wang, J.N.; Wang, Y.P.; Wang, H.C.; Tang, B.Y. Atomic diffusion mediated by vacancy defects in pure and transition element (TM)-doped (TM = Ti, Y, Zr or Hf) L12-Al3Sc. Mater. Des. 2016, 108, 529–537. [Google Scholar] [CrossRef]
  24. Henkelman, G.; Uberuaga, B.P.; Jónsson, H. A climbing image nudged elastic band method for finding saddle points and minimum energy paths. J. Chem. Phys. 2000, 113, 9901–9904. [Google Scholar] [CrossRef]
  25. Schmidt, C.; Bocquet, J. Calculation of the diffusion parameters in an ordered Ni3Alalloy for a relaxed lattice. Proc. Mater. Res. Soc. 1998, 527, 165–170. [Google Scholar] [CrossRef]
  26. Pan, S.; Li, C.; Qian, F.; Hao, L.; Li, Y. Diffusion controlled early-stage L12-D023 transitions within Al3Zr. Scr. Mater. 2023, 231, 115460. [Google Scholar] [CrossRef]
  27. Zhang, C.; Jiang, Y.; Cao, F.; Hu, T.; Wang, Y.; Yin, D. Formation of coherent, core-shelled nano-particles in dilute Al-Sc-Zr alloys from the first-principles. J. Mater. Sci. Technol. 2019, 35, 930–938. [Google Scholar] [CrossRef]
  28. Nityananda, R.; Hohenberg, P.; Kohn, W. Inhomogeneous electron gas. Resonance 2017, 22, 809–811. [Google Scholar] [CrossRef]
  29. Kresse, G.; Furthmüller, J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comp. Mater. Sci. 1996, 6, 15–50. [Google Scholar] [CrossRef]
  30. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758–1775. [Google Scholar] [CrossRef]
  31. Liu, X.; Wang, Q.; Zhao, C.; Li, H.; Wang, M.; Chen, D.; Wang, H. Formation of ordered precipitates in Al-Sc-Er-(Si/Zr) alloy from first-principles study. J. Rare Earths 2023, 9, 609–620. [Google Scholar] [CrossRef]
Figure 1. Formation energies of vacancy and antisite defects of L12-Al3M phase: (a) Al3Sc defect; (b) Al3Zr defect; (c) Al3Er defect; and (d) Al3Y defect.
Figure 1. Formation energies of vacancy and antisite defects of L12-Al3M phase: (a) Al3Sc defect; (b) Al3Zr defect; (c) Al3Er defect; and (d) Al3Y defect.
Molecules 28 06727 g001
Figure 2. Energy profiles for Al atom diffusion in Al3M phase: (a) NNJ(Al-VAl); (b) NNNJ1(Al-VAl); (c) NNNJ2(Al-VAl); (d) NNJ(Al-VM); and (e) ASB(AlM-Al-VM). The blue arrow represented the path of atomic jump.
Figure 2. Energy profiles for Al atom diffusion in Al3M phase: (a) NNJ(Al-VAl); (b) NNNJ1(Al-VAl); (c) NNNJ2(Al-VAl); (d) NNJ(Al-VM); and (e) ASB(AlM-Al-VM). The blue arrow represented the path of atomic jump.
Molecules 28 06727 g002
Figure 3. Energy profiles for M (M = Sc, Zr, Er, Y) atomic diffusion in Al3M phase: (a) NNJ(M-VAl); (b) AS(MAl-VAl); (c) ASB(MAl-M-VAl); (d) straight 6JC; (e) bent 6JC; and (f) NNNJ(M-VM). The blue arrow represented the path of atomic jump.
Figure 3. Energy profiles for M (M = Sc, Zr, Er, Y) atomic diffusion in Al3M phase: (a) NNJ(M-VAl); (b) AS(MAl-VAl); (c) ASB(MAl-M-VAl); (d) straight 6JC; (e) bent 6JC; and (f) NNNJ(M-VM). The blue arrow represented the path of atomic jump.
Molecules 28 06727 g003
Figure 4. Atomic-migration energy of core-shelled L12-Al3Zr(N): (a) the migration of Zr atom of Al3Zr shell layer to core Al3N phase and (b) the migration of N atom of Al3N core layer to shell Al3Zr phase. The blue arrow represented the path of atomic jump.
Figure 4. Atomic-migration energy of core-shelled L12-Al3Zr(N): (a) the migration of Zr atom of Al3Zr shell layer to core Al3N phase and (b) the migration of N atom of Al3N core layer to shell Al3Zr phase. The blue arrow represented the path of atomic jump.
Molecules 28 06727 g004
Table 1. Diffusion activation barrier of L12-Al3M.
Table 1. Diffusion activation barrier of L12-Al3M.
SystemDiffusion
Atoms
Diffusion
Mechanisms
Activation Barrier (eV)
Al-RichSc-Rich
Al3ScAlNNJ(Al-VAl)2.131–2.250 [23]1.778–1.689 [23]
NNJ(Al-VSc)1.668–2.352 [23]3.081–3.916 [23]
NNNJ1(Al-VAl)3.503–3.825 [23]3.150–3.244 [23]
NNNJ2(Al-VAl)5.963–6.441 [23]5.609–5.860 [23]
ASB(AlSc-Al-VSc)2.346–2.628 [23]5.574–6.297 [23]
ScNNJ(Sc-VAl)2.900–3.124 [23]2.547–2.543 [23]
AS(ScAl-VAl)3.956–4.144 [23]1.178–1.148 [23]
ASB(ScAl-Sc-VAl)5.079–5.546 [23]2.958–2.640 [23]
NNNJ(Sc-VSc)5.893–8.881 [23]7.306–10.625 [23]
S6JC(Sc----VAl)5.489–4.759 [23]5.135–6.504 [23]
B6JC(Sc----VAl)9.430–10.154 [23]9.076–10.734 [23]
Al3ZrAlNNJ(Al-VAl)1.7761.572
NNJ(Al-VZr)1.1751.788
NNNJ1(Al-VAl)3.3763.172
NNNJ2(Al-VAl)6.1835.830
ASB(AlZr-Al-VZr)1.7363.166
ZrNNJ(Zr-VAl)2.9262.721
AS(ZrAl-VAl)--
ASB(ZrAl-Zr-VAl)4.6663.645
NNNJ(Zr-VZr)6.0296.642
S6JC(Zr----VAl)5.3735.169
B6JC(Zr----VAl)8.7558.550
Al3ErAlNNJ(Al-VAl)1.8561.722
NNJ(Al-VEr)1.7482.149
NNNJ1(Al-VAl)1.9691.888
NNNJ2(Al-VAl)6.1876.053
ASB(AlEr-Al-VEr)2.8495.932
ErNNJ(Er-VAl)2.9662.832
AS(ErAl-VAl)--
ASB(ErAl-Er-VAl)5.3702.555
NNNJ(Er-VZr)6.4516.853
S6JC(Er----VAl)5.3505.216
B6JC(Er----VAl)9.2209.086
Al3YAlNNJ(Al-VAl)1.7301.649
NNJ(Al-VY)1.7251.968
NNNJ1(Al-VAl)1.9691.888
NNNJ2(Al-VAl)6.0906.009
ASB(AlY-Al-VY)2.8853.452
YNNJ(Y-VAl)4.5074.426
AS(YAl-VAl)--
ASB(YAl-Y-VAl)5.3474.942
NNNJ(Y-VSc)6.4726.715
S6JC(Y----VAl)6.2036.172
B6JC(Y----VAl)10.23010.149
Table 2. Formation energy of VAl defect in pure and Zr-substituted Al3N core.
Table 2. Formation energy of VAl defect in pure and Zr-substituted Al3N core.
Formation Energy of VAl (eV)
Al3ScAl3ErAl3Y
Pure0.863–1.2171.074–1.2081.026–1.107
Zr substitution1.1941.1871.063
Table 3. Formation energy of VAl defect in pure and N-substituted Al3Zr shell.
Table 3. Formation energy of VAl defect in pure and N-substituted Al3Zr shell.
PureScErY
VAl (eV)0.661–0.8650.8841.1361.359
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

Liu, S.; Liao, B.; Nie, B.; Fan, T.; Chen, D.; Zhang, J.; Song, Y. First-Principles Study of Atomic Diffusion by Vacancy Defect of the L12-Al3M (M = Sc, Zr, Er, Y) Phase. Molecules 2023, 28, 6727. https://doi.org/10.3390/molecules28186727

AMA Style

Liu S, Liao B, Nie B, Fan T, Chen D, Zhang J, Song Y. First-Principles Study of Atomic Diffusion by Vacancy Defect of the L12-Al3M (M = Sc, Zr, Er, Y) Phase. Molecules. 2023; 28(18):6727. https://doi.org/10.3390/molecules28186727

Chicago/Turabian Style

Liu, Shuai, Binbin Liao, Baohua Nie, Touwen Fan, Dongchu Chen, Jianglong Zhang, and Yu Song. 2023. "First-Principles Study of Atomic Diffusion by Vacancy Defect of the L12-Al3M (M = Sc, Zr, Er, Y) Phase" Molecules 28, no. 18: 6727. https://doi.org/10.3390/molecules28186727

Article Metrics

Back to TopTop