Next Article in Journal
BN Diamane-like Quasicrystal Based on 30° Twisted H-BN Bilayers and Its Approximants: Features of the Atomic Structure and Electronic Properties
Previous Article in Journal
On the Oscillating Course of dhkl−sin2ψ Plots for Plastically Deformed, Cold-Rolled Ferritic and Duplex Stainless Steel Sheets
Previous Article in Special Issue
First-Principles Study of the Structural, Mechanical and Thermodynamic Properties of Al11RE3 in Aluminum Alloys
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

First-Principle Investigation of the Interface Properties of the Core-Shelled L12-Al3M (M = Sc, Zr, Er, Y) Phase

1
School of Materials Science and Hydrogen Energy, Foshan University, Foshan 528000, China
2
Guangdong Key Laboratory for Hydrogen Energy Technologies, Foshan 528000, China
3
School of Mechatronic Engineering and Automation, Foshan University, Foshan 528000, China
4
College of Science, Hunan Institute of Technology, Hengyang 421002, China
*
Authors to whom correspondence should be addressed.
Crystals 2023, 13(3), 420; https://doi.org/10.3390/cryst13030420
Submission received: 22 January 2023 / Revised: 17 February 2023 / Accepted: 25 February 2023 / Published: 28 February 2023
(This article belongs to the Special Issue Advances in High Strength Steels)

Abstract

:
The interface structure and segregation behavior of L12-Al3M (M = Sc, Zr, Er, Y) phases were investigated based on first-principles calculations. The results showed that the order of the thermodynamically stable interface was Al3Zr/Al > Al3Sc/Al > Al3Er/ Al > Al3Y/Al. The interfaces of Al3Sc/Al3Zr, Al3Er/Al3Zr, and Al3Y/Al3Er obtained negative interfacial energies and low coherent strain energies and were favorable to form a clear interface. Zr atom tended to segregate to the first atomic layer on the Al side of the Al/Al3Sc, Al/Al3Er, and Al/Al3Y interfaces. The driving effect of the Zr atom segregation to the Al3Y shows was stronger than that to Sc and Er atoms, whereas the high coherent strain energy hindered the formation of Al3Y/Al3Zr interface. Er atom tended to segregate at the Al/Al3Y interface and accelerated the formation of core-shelled Al3Y/Al3Er. Furthermore, the formation of the double core-shelled Al3Y/Al3Er/Al3Zr was discussed.

1. Introduction

Aluminum alloys containing scandium had attracted the attention of researchers due to their high strength, corrosion resistance, and weldability [1,2,3]. The small addition of Sc to aluminum alloy formed a coherent Al3Sc phase (L12 phase) [4]. The primary Al3Sc phase can refine the grain of the as-cast Al alloy. The secondary Al3Sc phase effectively inhibited the growth of recrystallized grains [5] and increased the recrystallization temperature [6]. However, Al3Sc precipitate coarsened rapidly under high temperatures, reducing the strength of Al-Sc alloys. Moreover, the high price of Sc is prohibited from expanding commercial applications containing Sc aluminum alloys.
It was necessary to investigate new alloy elements to replace Sc elements and develop low-cost, high-performance aluminum alloys. Harada et al. [7] found that the atoms that replaced Sc to form the L12 microstructure should be located near Sc in the periodic table. Zr, Er, Yb, Y, and other elements may be good substitutes for Sc. Compared with the addition of Sc, the combined addition of Sc and Zr can precipitate a core-shelled Al3(Sc, Zr) phase with a Sc-rich core and Zr-rich shell microstructure owing to the differences in diffusion rates between Sc and Zr atoms. The core-shelled Al3(Sc, Zr) obtained better high-temperature stability (slower coarsening rate) [8]. The research of Er element alloying had attracted much attention due to the formation of L12-Al3Er phase and its low cost. By partially replacing the Sc element with Er element, the alloy can precipitate core-shelled Al3(Sc, Er) nano-phase with an Er-rich core and Sc-rich shell microstructure during homogenization [9], improving the creep resistance owing to its high lattice mismatch. Nie et al. [10] developed Al-Er alloys by completely replacing Sc with Er, and the alloys can also precipitate core-shelled Al3(Er, Zr) phase similar to Al3(Sc, Zr), displaying excellent thermal stability. Addition of Zr and Yb atoms into aluminum alloy promoted to precipitate L12-Al3(Zr, Yb) nano-phase. The core-shelled microstructure was observed in Al-Yb-Zr alloys as the Zr atom was incorporated into the Al3Yb nano-phase, thus displaying better thermal stability and inhibit-recrystallization ability [11]. The formation of the L12 nano-phase with core-shelled microstructure was usually attributed to the difference in element diffusion rate. The rapidly diffusing elements were enriched to form a core layer, and the slowly diffusing elements were segregated to form a shell layer.
Y and Sc were the congeners of the elements, and they had similar chemical properties. Zhang et al. [12] showed that Al-Y-Zr alloy can form L12-Al3(Zr, Y) precipitate at 400 °C. The Al3Y phase was first precipitated in the Al-Zr-Y alloy at the initial aging stage, which became the heterogeneous core of Al3Zr precipitated by aging, accelerating the precipitation of Zr in solid solution in the alloy. Based on the atomic diffusion control mechanism, Al3(Y, Zr) nano-phase should form a core-shelled microstructure with a Y-rich core and Zr-rich shell. However, Y and Zr atoms were evenly distributed in the precipitated phase, and no obvious segregation in the core/shell was observed in Al3(Y, Zr) phase [13]. The mechanism of the L12-Al3(Zr, Y) phase with no core-shelled microstructure had not been clearly understood.
The complex multilayer core-shelled microstructure provided a new idea for the structure design of nano-phase Al alloys. Christian Monachon et al. [14] performed a two-step aging treatment on the microalloyed Li, Sc, and Yb aluminum alloys, and high-density α-Al3(Li0.57Sc0.33Yb0.10) nano-sized precipitates were formed during aging treatment at 325 °C, exhibiting a Yb-enriched core and a Sc-enriched first shell with good stability due to the diffusivity ordering of DYb > DSc. During the second aging treatment at 170 °C, a second metastable δ’-Al3Li shell presented around the Sc-enriched first shell, thus illustrating an Al3Yb/Al3Sc/Al3Li bilayer core-shelled microstructure. Multilayer core-shelled microstructures had been observed in Al-Sc-Er-Zr alloys. After aging at 400 °C, Al3Er/Al3Sc/Al3Zr precipitates containing an Er-enriched core surrounded by a Sc-enriched inner shell and a Zr-enriched outer shell were formed in Al-Zr-Sc-Er alloys, obtaining a high coarsening resistant and high strength [15]. The formation of the bilayer core-shelled microstructure was well agreed with the atomic diffusivity ordering of DEr > DSc > DZr. However, the Al3Zr/Al(Yb, Sc) phase of the Zr core layer and shell layer with randomly distributed Yb and Sc was observed in Al-Yb-Sc-Zr alloys [16], and there was no Al3Yb/Al3Sc/Al3Zr bilayer core-shelled microstructure formed by the diffusion rate ordering of DYb > DSc > DZr. It implied that the diffusion rate did not control the formation of multilayer core-shelled microstructures.
First-principles thermodynamic calculation was an important method for understanding and predicting the structure and properties of the nanometer phases [17]. Based on first-principle thermodynamic calculations, Jiang et al. [18,19,20,21] calculated the interface formation energies and interfacial coherent strains on L12 nano-phases for Al-Sc/Er-Zr, Al-Sc-Er, and Al-Sc-Er-Zr alloys. These investigations showed that Sc, Er, or Zr strongly preferred to substitute the X sublattice sites in L12-Al3X (X = Er or Sc), while the inter-substitution between L12-Al3X was only weakly feasible. The (100)/(100) contacting facets were the most energy favored for the Al/Al3X, Al/Al3Zr, and Al3X/Al3 Zrinterfaces, and these interfaces were coherent with low formation energy. Zr segregation to the Al/Al3X interface promoted the formation of a Zr-enriched shell. Thus, L12-Al3(Rex, Zr1−x) precipitates tended to form the core-shelled microstructure.
As for L12-Al3(Zr, Y) precipitate, there was few first-principles calculation of interface energy. Li et al. [22] investigated the interactions at the Al3Y/Al interfaces based on first-principles calculation. However, first-principle calculations of the interfacial properties of multicomponent L12-phases containing Y atoms were rarely reported. In this paper, the interfacial properties of the Al3M(Sc, Er, Zr, Y) phase were investigated based on first-principles calculations to reveal the mechanism of the L12-Al3(Zr, Y) phase with no core-shelled microstructure. Furthermore, the complex multilayer core-shelled microstructure containing the Y atom was also discussed in this paper.

2. Computational Method

2.1. Calculation Details

This work was based on density functional theory (DFT) [23] and used VASP software [24] and the projection augmented wave (PAW) method [25] for first-principles calculations. The plane wave base cutoff energy used for structural optimization was 400 eV. The electron configuration was described by 3s23p1, 3s23p64s13d2, 4s24p65s14d3, 6s25p65d1, and 4s24p65s14d2 valence states for Al, Sc, Zr, Er, and Y, respectively. The Perdew-Burke-Ernzer (PBE) [26,27] method of generalized gradient approximation (GGA) was used to describe the exchange-correlation energy functional between electrons. The total energy was calculated using the linear tetrahedron method with the Blöchl correction until the total energy converged to 10−4 eV/atom. In each periodic direction of reciprocal space, the Monkhorst-Pack k point grids with linear kmesh analytical values of less than 0.03 2π Å−1 were used to integrate the Brilloin region, optimize the geometric structure, and calculate the follow-up result. The sandwich model with two interfaces without vacuum and the full optimization scheme including relaxation of atomic position, shape, and volume of the computational cell were used.

2.2. Interface Energy

To determine the appropriate interface structure, the interfacial energy and the coherent strain energy between the phase of L12-Al3M (M = Sc, Zr, Er, Y) and Al matrix were calculated, respectively. To evaluate the formation of the different interfacial, the interface formation energy ΔGf was calculated as follows [28]:
Δ G f = E Al Al 3 M i n t
where Eint was the total energy of the interface model; x denoted the phase fraction of Al; N was the total number of atoms in this interface model; and μAl and μAl3M were the fully relaxed energy per atom in Al and Al3M structures, respectively.
The interface formation energy contained two parts of energy, one was the contribution of interfacial energy, and the other was the elastic strain energy from the lattice mismatch, so it can also be expressed as follows [29]:
Δ G f N = 2 A γ N + G S
where A was the interface area. Since the calculation used a sandwich model with two interfaces (Figure 1), it was multiplied by a factor of two. Gs was the coherent strain energy per atom. γ was the interface energy per unit area without coherent strain energy.
In this calculation, the direct calculation method was used to calculate the energy of Al and Al3M phases after the relaxation in the direction perpendicular to the interface by fixing the lattice, and the difference was made with the total energy of the interface supercell. The interface energy γ can be calculated by the following formula [30]:
γ = 1 2 A [ E N Al N t o t a l E Al fix N Al 3 M N t o t a l E Al 3 M fix i n t [ ] ]
where NAl and NAl3M were the numbers of Al and Al3M phases, respectively; Ntotal was the total number of phases in the interface model; and E Al fix and E Al 3 M fix were the energies of Al and Al3M after the full relaxation perpendicular to the interface.
Based on Equation (3), the interface energy without the coherent strain can be obtained. The coherent strain energy can be calculated by applying Equations (1) and (2).
According to the research [18], the most energy-favored interface structures were determined as the Al-terminated and bridge-coordinated interface for the (100)Al/(100)Al3M, the Al-terminated and hollow-coordinated interface for the (110)Al/(110)Al3M, and the Al3M-terminated and hollow-coordinated interface for the (111)Al/(111)Al3M, respectively. The Al and Al3M block consisted of more than five layers. After full relaxation of the interface to obtain the total energy of the interface Eint, the lattices parallel to the interface a and b direction were fixed, and the interface model was replaced with pure Al or pure Al3M block of the same size, then relaxing in direction c to obtain E Al fix and E Al 3 M fix . These models were employed for the calculation in this work, and the Al/Al3M/Al interface models were shown in Figure 1. Directions a and b represent the interface plane, while direction c is perpendicular to the interface direction.
In order to analyze the formation of the core-shell structure from the perspective of thermodynamics, the interfacial energy and the coherent strain energy between the different L12 phases were also calculated, and the models were adopted as shown in Figure 2.

2.3. Interface Segregation Energy

The interfacial segregation of atoms had a thermodynamic driving force to take the place of the original atoms on the interface, thus forming a new interfacial structure [18,19]. In order to investigate the formation mechanism of the core-shelled L12 phase, the segregation energy of atoms at various atomic layers near the different Al3M interfaces was further investigated, which was calculated as the total energy difference before and after N atom segregation to the Al/Al3M interfaces. The segregation energy was expressed as [31]:
E s e g Al 3 M / N = ( E inter Al 3 M / N + E bulk site ) ( E inter Al 3 M + E bulk N )
where Al3M/N was the segregation of N atom from matrix to the Al3M interfaces; E inter Al 3 M / N was the total energy of the interface supercell after substitution by N atom in different layer or site; E inter Al 3 M was the total energy of the original interface supercell; and E bulk site and E bulk N were the atom chemical potential (site = Al or M) in the Al solid solution calculated using a 3 × 3 × 3 fcc supercell model.

3. Results and Discussion

3.1. Interface Energy

The direct calculation method was used to separate the interfacial energy and the coherent strain energy of Al3M/Al and Al3M/Al3N. The calculation results were shown in Table 1 and Table 2. Table 1 illustrated the interface energy and strain energy of the Al3M/Al interface in the (100), (110), and (111) facets. The calculation values of interface energy and strain energy for the Al3Er/Al, Al3Sc/Al, and Al3Zr/Al interfaces generally agreed with those in other literature [18,19] using linear fittings. For the Al3Y/Al interface, the calculations of interface energy were lower than that of the investigation by Li [22], which was related to the difference between the calculation methods. However, the order of the interface energy of different crystal facets was consistent with the reference [22], that was, Al(111)/Al3Y(111) > Al(100)/Al3Y(100) > Al(110)/Al3Y(110).
The order of the strain energy of the interface between the L12 phase and the Al matrix was Al3Y > Al3Er > Al3Sc > Al3Zr. However, the calculation values of strain energy in (110) facets for the Al3Er/Al, Al3Sc/Al, and Al3Zr/Al interfaces were higher than those in other literature [18,19] using linear fittings. In particular, the coherent strain energy of the Al3Y/Al and Al3Er/Al interfaces was much higher than that of the Al3Sc and Al3Zr phases. The coherent strain energy of L12-Al3M was related to the size mismatch in the lattice constant. Compared with the pure Al lattice constant 4.049 Å, the order of the size mismatch in the lattice constant was Al3Y > Al3Er > Al3Sc > Al3Zr (Table 1), which was well agreed with the order of the strain energy of the interface. Moreover, it can be inferred that Al3Er and Al3Y nano-phases obtained higher creep resistance than Al3Sc and Al3Zr nano-phases due to their higher coherent strain energy. The effect of the Er element on the creep resistance of aluminum alloy has been verified experimentally [32].
The order of the interface energy between the L12 phase and the aluminum interface was Al3Y > Al3Er > Al3Sc > Al3Zr. The results indicated that the order of formation for the thermodynamically stable interface was Al3Zr > Al3Sc > Al3Er > Al3Y. The low interface energy of Al3M/Al implied the thermodynamic driving force for the formation of a stable L12-Al3M nano-phases when the L12-Al3M nano-phases precipitated from the Al matrix. The lowest interface energy of Al3Zr indicated the low coarsening rate of the Al3Zr phase. L12-Al3Sc and Al3Er nano-phases were thermodynamically stable in aluminum alloys containing Sc, Er elements, which was well agreed with the previous investigation [6,10]. L12-Al3Y nano-phase was reported to form in the rapidly solidifying Al-Y alloys during solidification [33].
Furthermore, the crystal facets had a significant effect on the interface energy of Al3M/Al. The interface energy of Al(100)/Al3Sc(100) was the lowest, indicating that the Al3Sc phase tended to precipitate at (100) facets. The interface energy of (100) and (111) in Al3Zr was lower than that of the (110) facets, indicating that (100) and (111) facets were thermodynamically favored interfaces. Al3Er was inclined to precipitate in (100) and (110) facets. The calculation results showed that the (110) facets of Al3Y/Al were predicted as more energy favored than (100) and (111) facets.
Al3Zr was considered to be the ideal shell for thermodynamically stable nano precipitates due to the low coarsening rate of the Al3Zr phase and the low diffusion rate of the Zr element. However, the L12-Al3Zr nano-phase was thermodynamically metastable and transformed into the equilibrium D023 structure above 475 °C [34]. It was expected to improve the thermodynamic stability of the L12-Al3Zr nano-phase through Sc, Er elements microalloying. The interface energy and strain energy between L12-Al3M (Sc, Er, Y)/ Al3Zr in the (100), (110), and (111) facets were shown in Table 2. The calculation of L12-Al3Sc/Al3Zr and L12-Al3Er/Al3Zr were generally consistent with other literature [18,19] using linear fittings. The interface energies between different L12 phases were very small negative (essentially zero), and it indicated that such hetero-interfaces were thermodynamically stable.
However, the interface of L12-Al3Y/Al3Zr obtained the highest coherent strain energy in comparison with other L12 phase interfaces. It was difficult to form a good coherent interfacial relationship. Previous studies [12] showed that Al3Y was first precipitated in Al-Y-Zr alloy at the early aging stage and accelerated the precipitation of Zr. However, after long aging, the precipitated phase failed to form an obvious Al3Y core/Al3Zr shell microstructure, which agreed with the calculation results. On the other hand, the interface of L12-Al3Sc/Al3Zr and L12-Al3Er/Al3Zr had lower interface energy and better interfacial structural relationships than that of Al3Y/Al3Zr, thus presenting a stable core-shelled microstructure [18,19]. It can be inferred that core-shelled Al3Er/Al3Zr nano-phases obtained higher creep resistances than Al3Sc/Al3Zr and Al3Er/Al3Y nano-phases due to their higher coherent strain energy. Furthermore, the Al3Er/Al3Y and Al3Er/Al3Zr phases tended to form the core-shelled microstructure due to their low coherent strain energy and negative interfacial energy. If the Al3Er interlayer can be formed between Al3Y/Al3Zr, the stability of the double core-shelled Al3Y/Al3Er/Al3Zr can be improved in comparison with that of Al3Y/Al3Zr.
On the other hand, the (100) facet was the most thermodynamic favored interface among the (100), (100), and (100) facets due to its low interface energy and strain energy. The interface energy of Al3M(100)/Al3N(100) was the lowest, indicating that L12-Al3M (Sc, Er, Y)/Al3Zr and L12-Al3Er/Al3Y phases tended to precipitate at (100) facets. The investigation by Wang [35] also revealed that Al3Y(001)/Al(001) was the most stable interface. Moreover, L12-Al3M (Sc, Er, Y)/Al3Zr and L12-Al3Er/Al3Y phases maintained a good coherent interface at (100) facets owing to their lowest strain energy.

3.2. Segregation Energy

In order to clarify the formation mechanism of core-shell structure and the tendency of the second solute atom segregation to the interface, the segregation behavior of the second solute element on the interface was further investigated based on the interface calculation. The thermodynamic driving forces of atomic segregation at the interface were generally estimated by Equation (4).
The segregation energy of Zr atoms at various atomic layers near the different Al3M (Zr, Sc, Er, Y)/Al interfaces were shown in Figure 3. The Zr atom preferentially occupied the first layer of the interface at the Al matrix side. Except for plane (100), the segregation energy at Al was greater than that at M, indicating that Zr always tended to be biased towards the M site in the first layer (the first layer) on the Al side. In particular, for the Al3Sc(100)/Al(100) facet, there was a negative segregation energy for Zr occupying the Al position, indicating that Zr may precipitate at the Al position on the surface. Except for the (100) plane, Zr atoms did not occupy the Al position. It can be seen from the above results that Zr tended to form coherent structures during the segregation of the Al3Sc/Al, Al3Er/Al, and Al3Y/Al interface.
In addition, it can be seen from Figure 3d, the segregation behavior of Er atom towards the Al3Y/Al interface was similar to that of the Zr atom, suggesting that the mixture of Y and Er may be similar to Al3(Sc, Zr) and Al3(Er, Zr) to form a coherent interface of Al3Y/Al3Er.
The difference between the segregation energy at the interface and the energy dissolving in the Al matrix or Al3M phase represented the driving ability of the atom to segregate at the Al3M/Al interface. In order to see the numerical difference of segregation energy in Figure 3, the segregation energy of the Al3M layer (layer −5) and Al layer (layer 5) was subtracted, respectively, and the interface (layer 1) obtained ΔEAl3M and ΔEAl. The results are shown in Table 3. ΔEAl was the segregation energy difference to migrate from the interface to the Al matrix. The larger the ΔEAl, the greater the driving effect of the matrix-dissolved atoms’ segregation at the interface. ΔEAl3M was the segregation energy to be overcome when atoms migrated from the interface into the Al3M phase. The larger the value of ΔEAl3M, the more difficult it was for atoms to migrate into the Al3M phase.
Similarly, Table 3 showed that the ΔEAl3M and ΔEAl of the Zr atom were positive, indicating that the Zr atom was inclined to segregate at the Al/Al3Sc, Al/Al3Er, and Al/Al3Y interfaces, and the interfacial segregation of Zr accelerated the formation of core-shelled microstructures. On the other hand, the order of the ΔEAl3M and ΔEAl was Al3Y > Al3Er > Al3Sc. Al3Er and Al3Sc had a promoting effect on the segregation of the Zr atom. The Al3Sc/Al3Zr and Al3Er/Al3Zr interfaces can be formed owing to their small interfacial energy and strain energy. It indicated that, once the core-shelled microstructure was formed, it can stably maintain a clear core-shelled interface. The ΔEAl3M and ΔEAl of the Zr atom segregating to the Al/Al3Y interface were most positive, and the driving effect of the Y atom was stronger than that of Er, Sc atoms. The Y atom can promote the diffusion of the Zr atom, whereas it was difficult to form a stable core-shelled Al3Y/Al3Zr due to the large coherency strain energy and high mismatch between the Al3Y and Al3Zr. These calculation results were well consistent with the previous investigation [12,13].
Table 3 also showed that the ΔEAl3M and ΔEAl of Er atom to Al3Y were positive, and Er atom tended to segregate at the Al/Al3Y interface and accelerated the formation of core-shelled Al3Y/Al3Er. Al3Y had a precipitation-driving effect for Er atom. The diffusion rate of Er was higher than that of Zr, and the core-shelled Al3Er/Al3Zr can be formed. Therefore, as for Al-Y-Er-Zr alloys, they can obtain a thermodynamically stable Al3Y/Al3Er/Al3Zr with double core-shelled microstructure.

4. Conclusions

The interface structure and segregation behavior of L12-Al3M (M = Sc, Zr, Er, Y) phases were investigated based on first-principles calculations. The calculation results and conclusions were as follows:
(1)
The order of the thermodynamically stable interface was Al3Zr/Al > Al3Sc/Al > Al3Er/Al > Al3Y/Al. The interfaces of Al3Sc/Al3Zr, Al3Er/Al3Zr, and Al3Y/Al3Er obtained negative interfacial energies and low coherent strain energies and were favorable to form a clear interface. The high coherent strain energy hindered the formation of the Al3Y/Al3Zr interface.
(2)
Zr atoms tended to segregate to the first atomic layer on the Al side of the Al/Al3Sc, Al/Al3Er, and Al/Al3Y interfaces and occupied the Sc, Er, and Y lattice positions on the layer. The driving effect of the Zr atom segregation to the Al3Y was stronger than that of Al3Sc and Al3Er.
(3)
Er atoms tended to segregate at the Al/Al3Y interface and accelerated the formation of core-shelled Al3Y/Al3Er. It can obtain double core-shelled Al3Y/Al3Er/Al3Zr microstructure for Al-Y-Er-Zr alloys.

Author Contributions

B.N. and D.C. conceived and designed the research; S.L., Y.S., T.F., F.L. and H.Q. performed the first-principles calculation; Y.S. and S.Z. 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), the Science and Technology Program of the Ministry of Science and Technology (G2022030060L), Science and technology research project of Foshan (1920001000412, 2220001005305, 1920001001632), Science and technology project in Guangdong (2020b15120093, 2020B121202002), R and D plan for key areas in Jiangxi Province (20201BBE51009, 20212BBE51012), Foshan University Free Exploration Fund for graduate students (2021ZYTS10), Foshan Material Computing Engineering Technology Research Center and Innovation driven project of science and technology plan in Jiangxi Yichun.

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.

References

  1. Liu, L.; Xu, G.; Deng, Y.; Yu, Q.; Li, G.; Zhang, L.; Liu, B.; Fu, L.; Pan, Q. Existing form of Sc in metal-inert gas welded Al-0.60 Mg-0.75 Si alloy and its role in welding strength. Mater. Charact. 2023, 197, 112649. [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. Dorin, T.; Babaniaris, S.; Jiang, L.; Cassel, A.; Eggeman, A.; Robson, J. Precipitation sequence in Al-Sc-Zr alloys revisited. Materialia 2022, 26, 101608. [Google Scholar] [CrossRef]
  5. Ekaputra, C.; Weiss, D.; Mogonye, J.E.; Dunand, D.C. Eutectic, precipitation-strengthened alloy via laser fusion of blends of Al-7Ce-10Mg (wt.%), Zr, and Sc powders. Acta Mater. 2023, 246, 118676. [Google Scholar] [CrossRef]
  6. Zha, M.; Tian, T.; Jia, H.L.; Zhang, H.M.; Wang, H.Y. Sc/Zr ratio-dependent mechanisms of strength evolution and microstructural thermal stability of multi-scale hetero-structured Al–Mg–Sc–Zr alloys. J. Mater. Sci. Technol. 2023, 140, 67–78. [Google Scholar] [CrossRef]
  7. Harada, Y.; Dunand, D.C. Microstructure of Al3Sc with ternary transition-metal additions. Mater. Sci. Eng. A 2002, 329, 686–695. [Google Scholar] [CrossRef] [Green Version]
  8. Seidman, D.N.; Marquis, E.A.; Dunand, D.C. Precipitation strengthening at ambient and elevated temperatures of heat-treatable Al(Sc) alloys. Acta Mater. 2002, 50, 4021–4035. [Google Scholar] [CrossRef]
  9. 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 Earth. 2012, 30, 1276–1280. [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. 2022, 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. 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]
  13. 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]
  14. Christian, M.; David, C.D. Chemistry and structure of core/double-shell nanoscale precipitates in Al–6.5Li–0.07Sc–0.02Yb (at.%). Acta Mater. 2011, 59, 3398–3409. [Google Scholar]
  15. 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]
  16. 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]
  17. Zhao, Y.X.; Huang, Y.C.; Liu, Y. Insight into the stacking fault energy, dislocation, and thermodynamic properties of L12-Al3X (X = Sc, Ti, V) intermetallics from first-principles calculations. Mater. Today Commun. 2022, 31, 103684. [Google Scholar] [CrossRef]
  18. Zhang, C.; Jiang, Y.; Cao, F. 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]
  19. Zhang, C.; Yin, D.; Jiang, Y.; Wang, Y. Precipitation of L12-phase nano-particles in dilute Al-Er-Zr alloys from the first-principles. Comp. Mater. Sci. 2019, 162, 171–177. [Google Scholar] [CrossRef]
  20. Zhang, C.; Jiang, Y.; Guo, X. Formation and Relative Stabilities of Core-Shelled L12-Phase Nano-structures in Dilute Al–Sc–Er Alloys. Acta Metall. Sin.-Engl. 2020, 33, 1627–1634. [Google Scholar] [CrossRef]
  21. Zhang, C.M.; Xie, P.; Jiang, Y.; Zhan, S.; Ming, W.Q.; Chen, J.H.; Song, K.; Zhang, H. Double-Shelled L12 nano-structures in quaternary Al-Er-Sc-Zr alloys: Origin and critical significance. Acta Metall. Sin.-Engl. 2021, 34, 1277–1284. [Google Scholar] [CrossRef]
  22. Li, Y.; Huang, Y.; Zhang, X. Ab-initio studies of the micromechanics and interfacial behavior of Al3Y|fcc-Al. Metals 2022, 12, 1680. [Google Scholar] [CrossRef]
  23. Nityananda, R.; Hohenberg, P.; Kohn, W. Inhomogeneous electron gas. Resonance 2017, 22, 809–811. [Google Scholar] [CrossRef]
  24. 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]
  25. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758–1775. [Google Scholar] [CrossRef]
  26. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1998, 77, 3865–3868. [Google Scholar] [CrossRef] [Green Version]
  27. Budimir, M.; Damjanovic, D.; Setter, N. Piezoelectric Response and Free Energy Instability in the Perovskite Crystals BaTiO3, PbTiO3 and Pb(Zr, Ti)O3. Phys. Rev. B 2006, 73, 4106. [Google Scholar] [CrossRef] [Green Version]
  28. 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]
  29. Wang, Y.; Liu, Z.K.; Chen, L.Q.; Wolverton, C. First-principles calculations of β″-Mg5Si6/α-Al interfaces. Acta Mater. 2007, 55, 5934–5947. [Google Scholar] [CrossRef]
  30. Li, J.; Zhang, M.; Zhou, Y.; Chen, G. First-principles study of Al/A13Ti heterogeneous nucleation interface. Appl. Surf. Sci. 2014, 307, 593–600. [Google Scholar] [CrossRef]
  31. Scheiber, D.; Pippan, R.; Puschnig, P.; Romaner, L. Ab initio search for cohesion-enhancing impurity elements at grain boundaries in molybdenum and tungsten. Model. Simul. Mater. Sci. Eng. 2016, 24, 085009. [Google Scholar] [CrossRef]
  32. Vo, N.Q.; Dunand, D.C.; Seidman, D.N. Improving aging and creep resistance in a dilute Al–Sc alloy by microalloying with Si, Zr and Er. Acta Mater. 2014, 63, 73–85. [Google Scholar] [CrossRef]
  33. Foley, J.C.; Perepezko, J.H.; Skinner, D.J. Formation of metastable L12-Al3Y through rapid solidification processing. Mater. Sci. Eng. A 1994, 179, 205–209. [Google Scholar] [CrossRef]
  34. Knipling, K.E.; Dunand, D.C.; Seidman, D.N. Precipitation evolution in Al–Zr and Al–Zr–Ti alloys during aging at 450–600 °C. Acta Mater. 2008, 56, 1182–1195. [Google Scholar] [CrossRef]
  35. Wang, Y.; Miao, Y.; Peng, P.; Gao, H.; Wang, J.; Sun, B. Ab initio investigation on preferred orientation at the Al/Al3 (Zr, Y) interface in Al–Zr–Y alloy. J. Appl. Phys. 2022, 131, 225111. [Google Scholar] [CrossRef]
Figure 1. The interface model of Al/Al3M/Al: (a) Al(100)/Al3M(100)/Al(100); (b) Al(110)/Al3M(110)/Al(110); (c) Al(111)/Al3M(111)/Al(111).
Figure 1. The interface model of Al/Al3M/Al: (a) Al(100)/Al3M(100)/Al(100); (b) Al(110)/Al3M(110)/Al(110); (c) Al(111)/Al3M(111)/Al(111).
Crystals 13 00420 g001
Figure 2. The interface models of Al3N/Al3M/Al3N: (a) Al3N(100)/Al3M(100)/Al3N(100); (b) Al3N(110)/Al3M(110)/Al3N(110); (c) Al3N(111)/Al3M(111)/Al3N(111).
Figure 2. The interface models of Al3N/Al3M/Al3N: (a) Al3N(100)/Al3M(100)/Al3N(100); (b) Al3N(110)/Al3M(110)/Al3N(110); (c) Al3N(111)/Al3M(111)/Al3N(111).
Crystals 13 00420 g002
Figure 3. The segregation energies of Zr atom segregating into different atomic layers: (a) Al3Sc/Al; (b) Al3Er/Al; (c) Al3Y/Al. The segregation energies of Er atom segregating to (d) Al3Y/Al interface.
Figure 3. The segregation energies of Zr atom segregating into different atomic layers: (a) Al3Sc/Al; (b) Al3Er/Al; (c) Al3Y/Al. The segregation energies of Er atom segregating to (d) Al3Y/Al interface.
Crystals 13 00420 g003
Table 1. The lattice parameter interface energy and strain energy of Al3M/Al interface.
Table 1. The lattice parameter interface energy and strain energy of Al3M/Al interface.
InterfaceLattice Parameter (Å)Interface Energy γ (J/m2)Strain Energy GS (meV/Atom)
abc
Al/Al3Sc(100)4.07536.8840.1353.5
(110)4.0775.81834.6080.204 3.2
(111)5.76010.03935.2550.221 1.7
Al/Al3Zr(100)4.06137.1820.078 2.3
(110)4.0635.82434.6930.109 1.3
(111)5.76810.01535.5000.0761.3
Al/Al3Er(100)4.17737.0640.181 8.8
(110)4.1835.95535.1960.185 7.7
(111)5.93310.23935.4520.2279.0
Al/Al3Y(100)4.20137.0680.197 9.4
(110)4.2035.98635.3550.1859.6
(111)5.96410.28635.5130.23110.8
Table 2. The lattice parameter, interface energy, and strain energy of Al3M/Al3N interface.
Table 2. The lattice parameter, interface energy, and strain energy of Al3M/Al3N interface.
InterfaceLattice Parameter (Å)Interface Energy γ (J/m2)Strain Energy GS (meV/Atom)
abc
Al3Sc/Al3Zr(100)4.102-36.982−0.0790.1
(110)4.0995.79834.913−0.021 0.2
(111)5.80110.05035.579−0.056 0.2
Al3Er/Al3Zr(100)4.247-38.218−0.087 4.6
(110)4.2496.02835.868−0.047 6.5
(111)6.00710.39836.709−0.062 5.4
Al3Zr/Al3Y(100)4.171-37.554−0.0886.9
(110)4.1715.89935.511−0.0489.1
(111)5.89510.21636.112−0.0567.9
Al3Er/Al3Y(100)4.188-37.673−0.0180.1
(110)4.1875.92235.644−0.0040.2
(111)5.91510.25436.204−0.00040.2
Table 3. The difference of the segregation energy between interface and Al3M phase and Al layer of each L12 phase.
Table 3. The difference of the segregation energy between interface and Al3M phase and Al layer of each L12 phase.
Solute AtomsInterfaceΔEAl3M (eV/Atom)ΔEAl (eV/Atom)
ZrAl3Sc/Al(100)/(100)0.5520.464
(110)/(110)0.6530.459
(111)/(111)0.6440.457
Al3Er/Al(100)/(100)0.6150.540
(110)/(110)0.7490.449
(111)/(111)0.8100.617
Al3Y/Al (100)/(100)0.8980.530
(110)/(110)1.0860.488
(111)/(111)1.1310.653
ErAl3Y/Al(100)/(100)0.8070.354
(110)/(110)1.0300.480
(111)/(111)0.9910.659
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

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. https://doi.org/10.3390/cryst13030420

AMA Style

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(3):420. https://doi.org/10.3390/cryst13030420

Chicago/Turabian Style

Song, Yu, Songtao Zhan, Baohua Nie, Shuai Liu, Haiying Qi, Fangjun Liu, Touwen Fan, and Dongchu Chen. 2023. "First-Principle Investigation of the Interface Properties of the Core-Shelled L12-Al3M (M = Sc, Zr, Er, Y) Phase" Crystals 13, no. 3: 420. https://doi.org/10.3390/cryst13030420

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