Next Article in Journal
Crystallization of Intermetallic Phases Fe2Si, Fe5Si3 for High Alloyed Cast Irons
Previous Article in Journal
Preparation, Properties, and Applications of Near Stoichiometric Lithium Tantalate Crystals
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

First-Principles Investigation of the Diffusion of TM and the Nucleation and Growth of L12 Al3TM Particles in Al Alloys

1
School of Material Science and Hydrogen Energy Engineering, Foshan University, Foshan 528001, China
2
Guangdong Key Laboratory for Hydrogen Energy Technologies, Foshan 528000, China
3
School of Science and Research Institute of Automobile Parts Technology, Hunan Institute of Technology, Hengyang 421002, China
4
School of Mechatronic Engineering and Automation, Foshan University, Foshan 528001, China
*
Authors to whom correspondence should be addressed.
Crystals 2023, 13(7), 1032; https://doi.org/10.3390/cryst13071032
Submission received: 5 June 2023 / Revised: 26 June 2023 / Accepted: 27 June 2023 / Published: 29 June 2023
(This article belongs to the Topic Microstructure and Properties in Metals and Alloys)

Abstract

:
The key parameters of growth and nucleation of Al3TM particles (TM = Sc-Zn, Y-Cd and Hf-Hg) have been calculated using the combination of the first principles calculations with the quasi-harmonic approximation (QHA). Herein, the diffusion rate Ds of TM elements in Al is calculated using the diffusion activation energy Q, and the results show that the Ds of all impurity atoms increases logarithmically with the increase in temperature. With the increase in atomic number of TM, the Ds of 3–5d TM elements decreases linearly from Sc, Y and Hf to Mn, Ru and Ir, and then increases to Zn, Ag and Au, respectively. The interface energy γα/β, strain energy ΔEcs, chemical formation energy variation ΔGV and surface energy E s u r a v e were further computed from the based interface and slab models, respectively. It was found that, with the increase in the atomic number of TM, the interface energies γα/β of Al/Al3TM (TM = (Sc-Zn, Y-Cd)) decreased from Sc and Y to Mn and Tc and then increased to Zn and Cd, respectively (except for the (001) plane of Al/Al3(Fe-Co), the (111) plane of Al/Al3Pd and the (110) and (111) planes of Al/Al3Cd). The strain energies ΔEcs of Al/Al3TM (TM = (Sc-Zn)) increased at first, and then decreased for all cycles. The chemical formation energy ΔGV of all Al3TM changed slightly in the temperature range of 0~1000 K, except that the ΔGV of Al3Sc, Al3Cu, Al3(Y-Zr), Al3Cd, Al3Hf and Al3Hg increased nonlinearly. With the increase in atomic number at both 300 and 600 K, the ΔGV of 3–5d TM elements increased from Sc, Y and Hf to Mn, Tc and Re at first, and then decreased to Co, Rh and Ir, respectively, and slightly changes at the end. With the increase in atomic number of TM, the variation trends of the surface energies of Al3TM intermetallic compounds present similar changes for all cycles, and the (111) surface always has the lowest values.

1. Introduction

Al-based alloys have been widely applied in the electronics, aerospace and automotive industries due to their low density, high specific strength and welding strength [1]. Adding transition elements (TMs) can significantly improve the mechanical and thermodynamic properties of Al alloys [2,3,4,5,6]. For example, the existence of Sc (0.3%) in the Al matrix increases the ultimate rupture strength of annealed Al sheets from 55 to 240 MPa [7], and L12-Al3Zr in the Al matrix is used as a grain refiner to improve the coarsening resistance and creep properties [8,9]. However, the high cost of Sc and Zr limits their applications in commercial Al alloys. Specifically, intermetallic compounds with TMs are suitable candidates for high temperature applications, as the crystal types in the Al matrix may be L12, D022, D023 or D019 structures [8,10,11,12], of which the L12 phase is an important intermetallic compound and has been widely studied [13,14,15,16]. Moreover, the TMs can be used to substitute the expensive Sc and Zr elements in L12-Al3Sc and Al3Zr.
The previous research proved that fixing the dislocations and grain boundaries can effectively refine the deformed and recrystallized grains, depending on the dispersed distribution of L12 Al3TM particles during rising heat [17,18]. The diffusion rate of TM solute atoms in an Al matrix and the interfacial properties of Al3TM/Al are important parameters for the investigation of nucleation, the growth of L12 Al3TM phases [19,20,21], and the low-index bonds of particles to matrix [22,23]. However, the experimental exploration of appropriative substitution TMs is difficult because of the complex environment and the expensive cost [15,23,24,25,26,27,28]. Fortunately, in recent years, with the development of modern computer technologies, theoretical identification (e.g., first-principles (FP) calculations based on density functional theory (DFT) [15]) in the complicated systems (e.g., metals and ceramics) has become the most powerful method to accomplish this [29,30,31,32].
The stability and nucleation behavior of L12-Al3Sc and Al3Li binary phases have first been investigated using the framework of density functional theory (DFT) calculation by Mao et al. [15]. Their results showed that the L12-Al3Sc and Al3Li structures have lower formation energies than those of the corresponding D023, D019 and D022 structures. Furthermore, they found that the interface and strain energies of Al3Sc are much higher than those of Al3Li for all (001), (110) and (111) interfaces. Zhang et al. [33] have comprehensively studied the solubility of RE (RE = Y, Dy, Ho, Er, Tm and Lu) in Al based on the free energy difference between L12 bulk and Al solid solution matrix in the DFT theoretical framework. Their results indicated that the solubility of all rare earth (RE) (RE = Y, Dy, Ho, Er, Tm and Lu) elements increases with the increase in temperature (~1000 K). They also believed that Dy and Y elements can become better candidates for Sc due to the better stability of Al3Dy and Al3Y compounds and their almost identical solubility compared to the higher-cost Sc element. Sun et al. [34] have calculated low-index (001), (110) and (111) surface energies of L12-Al3Sc particles adopting slab model with 15 Å vacuum region. Their results show that when the surface energies of non-stoichiometric (001) and (110) surfaces of Al3Sc are calculated, their values should be considered as different under different Al chemical potentials, and in a wide range of Al chemical potentials, the surface energies of the (111) surface with AlSc-terminated have lower values, indicating that they are more stable than other surfaces.
However, up to date, the diffusion rates Ds of TMs in Al, the surface properties of L12-Al3TM and the interface of Al3TM/Al-matrix have not been systematically investigated. Specifically, the nucleation and growth of L12-Al3TM (TM = Sc-Zn, Y-Ag, Hf-Au) particles at finite temperatures have not been obtained, and their relationship to the atomic number of TM hasn’t been described in detail due to the large computational cost required. In the present work, by combining the first-principles calculations with the quasi-harmonic approximation (QHA), the relationship between the particles’ nucleation/growth and atomic number/temperature are discussed. First, the diffusion rates Ds of TMs as a functional of atomic number and temperature have been researched. Then, the relationship between the driving force and the hindrance of particle nucleation and the atomic number of TM is explained based on the interface model. Finally, the effects of the surface stability of different intermetallic compounds with the change in atomic number based on the slab model are obtained.

2. Computed Methods

All calculations in this work were performed in Vienna ab initio Simulation Package (VASP) [34] with the 5.4.4 version, which adopts the framework of density functional theory (DFT) [35] calculations to solve the Kohn–Sham equation and obtain the total energy from different models. In the calculated processing of VASP, to relax all models to their most stable ground state, the electron–core interaction was described by the projector augmented wave (PAW) [36] method. The optimal choice of exchange–correlation functional was considered using the generalized gradient approximation (GGA) with the Perdew–Burke–Ernzerh (PBE) version [31]. A 10 × 10 × 10 k-point sampling grid with the Gamma-centered Monkhorst–Pack method [37] in the first Brillouin zone was selected via strict convergence testing (see Figure 1) for bulk properties calculation. A cut off energy of the plane-wave basis of 500 eV was chosen for the whole calculated process. The energy and force tolerance were set to 10−7 eV and 0.01 eV/Å, respectively, by using conjugate gradient (CG) minimization and Broyden–Fletcher–Goldfarb–Shannon (BFGS) schemes [38].
Here, based on the slab model, we investigated three low-index surfaces, containing (100), (110) and (111) surfaces of Al and L12-Al3TM, which adopted 14, 14 and 16 layers, respectively [24,39]. All interfaces of (100), (110) and (111) surfaces of Al/L12–Al3TM are calculated by using the 18 layers interface model. A 10 × 10 × 1 k–mesh grid for both cases was tested to be suitable for this work. To simulate diffusion behavior, we constructed a 3 × 3 × 3 supercell with 4 × 4 × 4 k–mesh grids to obtain the diffusion barriers of the solute diffusion of TMs in the Al matrix based on the climbing-image nudged elastic band (CI-NEB) [40] method. Meanwhile, a spring force constant of 5 eV/Å was considered to keep all the images separated, and these CI-NEB iterations were continued until the forces on each atom were less than 0.05 eV/Å.

3. Conclusion Description

3.1. Diffusion

In the process of heating up, some atoms will detach from their original equilibrium positions and then diffuse to a new site while obtaining enough energy. Thus, the diffusion behavior is a common phenomenon in the field of material science and engineering. According to the Lifshitz and Slyozov and Wagner methods [41,42], the growth of particles is affected by the diffusion behavior of solute atoms, and the faster diffusion in the Al matrix is beneficial for the grain growth. In the current work, to investigate diffusion behavior, we first show a vacancy-substitution model, as depicted and visualized in Figure 2a by VESTA codes. The vacancy-substitution model can be divided into two types: the self-diffusion of the violet Al atom and the impurity diffusion of the TM pink ball [24,39]. The black arrow represents the diffusion path for the TM atom. To further investigate the diffusion behavior, the diffusion coefficient as a function of jump frequency I is expressed, which satisfies the Arrhenius equation as follows [40,43,44,45]:
D T = λ a 2 2 Z I
where λ (λ = 2), Z (Z = 1) and a are the number of directions for atomic transitions, the dimension of diffusion and the corresponding atomic distance of diffusion, respectively.
Here, the jump frequency for both diffusions in solid-state was established using the classical transition state theory (TST) [46,47]:
I = ν e x p Q κ T
where ν, Q, T and κ are the effective frequencies associated with the vibration of the transition atom, the diffusion activation energy, the special temperature and the Boltzmann constant, respectively.
According to Winter–Zener theory (WZT), the ν can be approximately expressed as [48]:
ν = 2 E D i f f m a 2 1 2
where m represents the atomic mass of transition atoms. Herein, two types of diffusion activation energies Q corresponding to self D0 and impurity Ds diffusion coefficients are gained using first principles calculations. The Q for self-diffusion contains two separate energies: vacancy formation energy Evac and the migration energy of Al atom Em in Al matrix. For impurity diffusion activation energy, the activation energy Q consists of three parts: the substitutional solution energy Es of a TM atom replacing a Al atom, vacancy formation energy Ef in the presence of TM in Al107TM supercell, and the migration energy of diffusion Em [49]:
E s = E A l 107 T M 107 E A l E T M
E f = E A l 106 T M : V a c E A l 107 T M + E A l
E b = E s + E f
where ETM and EAl are the energies of single TM and Al atoms in the stable bulk, respectively, and Eb is the binding energy of a TM atom substituting a vacancy in Al matrix.
To further investigate the physical mechanism of behaviors, the electron localized function (ELF) has been drawn using the VESTA code [50]. The ELF is defined as:
E L F = 1 1 + D r D h r 2
where D r and D h r are the true electron gas density and the pre-assumed uniform electron gas density, respectively.
The Em, Eb (Evac), Q and Ds (D0) for TM and Al at 300 K with available experimental and theoretical values are summarized in Table 1 [51,52,53,54]. It can be seen that errors between the present and previous values in literatures for Em, Eb (Evac) and EDiff are within 20%, and the current value of EDiff of Sc element is only ~2% larger than that of the experiment value. To visually illustrate the regularity of the variations of activation energy Q as a function of the atomic number of TM, it is further plotted in Figure 2b. The result shows that the Q increases at first and then decreases as the atomic number increases (Sc-Zn, Y-Ag, Hf-Au) in the Al matrix (except for Cr of 2.23 eV), indicating that there is a correlation between the valence electron configuration of impurity elements and the activation energy Q. Additionally, the TM elements in the fourth cycle generally have lower diffusion activation energies Qs, ranging from 0.35 to 2.60 eV. For Mn-Co, Tc-Rh and Re-Ir, they have larger Qs in the Al matrix, which are 2.45~2.60, 3.82~3.94 and 3.95~4.26 eV, respectively, indicating that their diffusion abilities are relatively weak in the Al matrix. Meanwhile, for Cu-Zn, Ag and Au, the activation energy Q is very low, or even negative for a Cd of −0.12 eV and an Hg of −0.30 eV, as shown in Table 1, which shows they are easier to move in the Al matrix. In the undoped-Al system, self-diffusion activation energies Q0 is lower compared to all Qs in the doped system, except for the Qs of Cu, Zn, Y and Ag, indicating that the diffusion of most TM atoms is more difficult than self-diffusion.
The variation in activation energy Q with the temperature increasing can be calculated from the above results by combining them with the quasi-harmonic vibration (QHA) [55]; by doing this, the change in diffusion rate D with the temperature can be obtained via Equations (1)–(3), and the results are summarized in Table 1 and Figure 2c,d. It should be noted that only the self-diffusion rate D0 as a function of Q is presented in the inlet of Figure 2c, owing to the fact that that all activation energies Q of TM elements are nearly the same. The self-diffusion rate D0 of 3.55 × 10−28 m2·s−1 for Al in this work is in general agreement with the experimental extension values from 1.76 × 10−27~4.42 × 10−12 m2·s−1 in the range of 300~1000 K and 1.47 × 10−14~1.36 × 10−12 in the range of 739~917 K in literature [56,57], seen from Table 1 and Figure 2c. Meanwhile, the theoretical predicted D0 of 3.55 × 10−28 m2·s−1 of Al is lower than that of the experiment at 300 K. The reason for this may be that it is difficult to accurately determine the D0 due to the influence of crystal structure defects, dislocations and grain boundaries in experiments. The Ds of all impurity atoms except for Cd and Hg increases logarithmically with the increase in temperature. A negative Q for Cd and Hg cases makes it impossible to theoretically calculate values according to Equations (1) and (2). Reasonably, the D indicate the inverse pattern to Q; higher barriers mean slower passage. Additionally, the larger the value at 300 K, the lower the increasing rate. This trend result is consistent with the variation trend of Ds for Mg, Si and Cu with temperature calculated by Mantina et al. [44]. Figure 2d further shows the diffusion rate Ds at 300 K as a function of the atomic number of TM, and it can be seen that the diffusion rate Ds first decreases linearly from 2.05 × 10−37, 6.47 × 10−24 and 2.79 × 10−44 m2·s−1 for Sc, Y and Hf to 2.43 × 10−50, 6.77 × 10−73 and 1.60 × 10−78 m2·s−1 for Mn, Ru and Ir and then increases with the increase in atomic number to 3.09 × 10−13, 9.17 × 10−17 and 2.93 × 10−29 m2·s−1 for Zn, Ag and Au, respectively (except for Cr of 3.36 × 10−44 m2·s−1).
From the above results, it can be seen that higher peaks occur for half- or near half-full d shells for all cycles considered. The reason for this may be that half- or near half-full d shells of the TM element in the Al matrix are more stable and more energy is required to force them to move from the stable site to the vacancy. Although the atomic diffusion barrier changes similarly with the increase in atomic number in the same period, TM with 3d shells present a faster diffusion behavior. To explore the underlying potential, the ELF of Sc and Ru doping systems on the (010) plane are presented in Figure 2e,f. The value of ELF, which is selected as 0 to 1, demonstrates the probability of finding an electron in the neighborhood space. To be specific, when it equals 0, it reflects a strongly delocalized electron area; when it equals 1, it corresponds to a strongly localized electron area. It can be seen that, when Sc and Ru are the first nearest neighbors of the vacancy, different values of ELF are exhibited. The Ru would make the surrounding electrons appear more likely than Sc, resulting in Ru being difficult to diffuse to the vacancy.
Figure 2. (a) The diffusion model. (b) The calculated diffusion barrier of a vacancy Em, vacancy solute binding energy Eb (vacancy formation energy Evac for self-diffusion in Al matrix) and diffusion activation energy EDiff with the change in atomic number. (c) The diffusion rate D and EDiff as a function of temperature. (d) The impurity diffusion rate Ds as a function of the atomic number of TM. (a,b) represent the experimental values from Murphy et al. [57] and Volin et al. [56], respectively. (e,f) The ELFs on the (010) planes of Sc and Ru doping systems, respectively.
Figure 2. (a) The diffusion model. (b) The calculated diffusion barrier of a vacancy Em, vacancy solute binding energy Eb (vacancy formation energy Evac for self-diffusion in Al matrix) and diffusion activation energy EDiff with the change in atomic number. (c) The diffusion rate D and EDiff as a function of temperature. (d) The impurity diffusion rate Ds as a function of the atomic number of TM. (a,b) represent the experimental values from Murphy et al. [57] and Volin et al. [56], respectively. (e,f) The ELFs on the (010) planes of Sc and Ru doping systems, respectively.
Crystals 13 01032 g002
Table 1. The calculated diffusion barrier Em (eV), vacancy–solute binding energy Eb (eV), diffusion activation energy Q (eV) and diffusion rate Ds (m2·s−1) for TM atoms in Al matrix at 300 K. It should be noted that for pure Al, Eb and Ds are in fact Evac and D0, respectively. Note: A negative activation energy Q can’t meet calculating Ds according to Equations (1)–(3).
Table 1. The calculated diffusion barrier Em (eV), vacancy–solute binding energy Eb (eV), diffusion activation energy Q (eV) and diffusion rate Ds (m2·s−1) for TM atoms in Al matrix at 300 K. It should be noted that for pure Al, Eb and Ds are in fact Evac and D0, respectively. Note: A negative activation energy Q can’t meet calculating Ds according to Equations (1)–(3).
ElementEmEbQDs
Al0.68
0.55–0.70 [54]
0.57 [58]
0.63
0.60–0.80 [54]
0.63 [58]
1.31
1.15–1.50 [54]
1.20 [58]
1.31 [56]
3.55 × 10−28
1.76 × 10−27 [56]
Sc0.85 0.97 1.82 1.79 [55]2.05 × 10−37
Ti1.43 0.84 2.27 6.04 × 10−45
V1.90 0.42 2.32 1.09 × 10−45
Cr2.14 0.09 2.23 3.36 × 10−44
Mn2.11 0.49 2.60 2.43 × 10−50
Fe1.90 0.63 2.53 3.10 × 10−49
Co1.55 0.90 2.45 7.79 × 10−48
Ni1.06 0.97 2.03 6.44 × 10−41
Cu0.57 0.22 0.79 2.94 × 10−20
Zn0.40 −0.04 0.35 3.09 × 10−13
Y0.36 0.63 0.99 6.47 × 10−24
Zr1.19 0.98 2.17 2.16 × 10−43
Nb1.88 0.75 2.63 4.94 × 10−51
Mo2.46 0.75 3.22 7.40 × 10−61
Tc2.54 1.27 3.82 7.25 × 10−71
Ru2.25 1.69 3.94 6.77 × 10−73
Rh1.68 2.16 3.84 2.81 × 10−71
Pd0.98 1.71 2.68 5.51 × 10−52
Ag0.51 0.06 0.56 9.17 × 10−17
Cd0.35 −0.46 −0.12 -
Hf1.41 0.80 2.21 2.79 × 10−44
Ta2.10 0.41 2.51 2.90 × 10−49
W2.85 0.17 3.02 9.43 × 10−58
Re3.09 0.86 3.95 2.93 × 10−73
Os2.77 1.36 4.14 1.96 × 10−76
Ir2.15 2.11 4.26 1.60 × 10−78
Pt1.27 2.15 3.42 1.77 × 10−64
Au0.53 0.78 1.31 2.93 × 10−29
Hg0.21 −0.50 −0.30 -

3.2. Nucleation

According to the classical nucleation method (CNT) [44,56,57], the total energy of the nucleation process of second phases can be expressed as follows: G t o t = 4 3 π R 3 G V + E C S + 4 π R 2 γ α / β . Here, a positive strain energy contribution would be a hindrance when Al3TM grains gradually form, while the difference in free energy in bulk between the matrix and particles and the interfacial free energy would promote particle nucleation.
Here, to calculate interface energy γα/β, we adopt a total energy of interface model that subtracts the total energy of the phases on either side of the interface in a two-phase system [23]:
γ α / β = E α / β E α + E β 2 A
where A is the area of the interface, Eα/β is the total internal energy of the relaxed α/β system containing an interface and Eα and Eβ are the total internal energies of phases α and β from the strains of all directions, respectively.
The chemical formation energy difference ΔGV of L12-Al3TM precipitates can be expressed in dilute solid solution based thermodynamics, AlnTMAl3TM + Aln−3. It can be shown as [15]
G V = G A l 3 T M + n 3 G A l G A l n T M
where n (n = 31) and ΔG are the number of atoms and Gibbs free energy, respectively. To investigate the dependence of ΔGV on temperature, the non-equilibrium free energy ΔGV is derived as the following equation [15,59]:
G V , P , T = m i n [ F V , T ] + P V
where F V ; T is the free energy computed by the sum of electronic internal energy and phonon Helmholtz free energy F V , T = U e l + F v i b . P is the circumstance pressure.
Due to lattice mismatch, both the harmonic and non-harmonic contributions were observed to calculate the strain energy ΔECS of the L12 precipitation phases [15]:
E C S x , Ĝ = min a s x E α e q i a s , Ĝ + 1 x E β e q i a s , Ĝ
where a s is the constrained superlattice parameter, Ĝ is the direction and x is the mole fraction of phase α. E α e q i and E β e q i are the epitaxial deformation energies of phases α and β, respectively.
Figure 3a shows the interface model for calculating the interface properties in this work. The Al matrixes are highlighted in dashed rectangles, and different layer numbers are used for the calculation convenience. Comparing the present results with references [15,23] listed in Table 2, there are larger errors compared by Mao and Li et al. [15,23], and these errors are further discussed. The main reasons are as follows:
  • Li et al. [23] adopted the vacuum slab model for the calculation, resulting in the values of interface energies being affected by different terminal surfaces, and the interface energy of Al/Al3Ti of 61.85 mJ·m−2 calculated by the vacuum model is in a good agreement with that of Li et al. according to γ α / β = E α / β * E s l a b , α + E s l a b , β S + E s u r α + E s u r β , where E α / β * is the total energy of the vacuum slab model system, E s l a b denotes the total energy of the fully relaxed surface slabs and E s u r α and E s u r β represent the surface energies of the α and β surface slabs, respectively. Meanwhile, the strain energy caused by lattice mismatch in the vacuum slab model was not taken into account in the above equation.
  • Mao et al. [15] had investigated interface properties in a periodic supercell and, considering the strain energy of interface model, they calculated interface properties with less accuracy, performed on a 0.13 (1/Å) spacing Monkhorst–Pack k-point mesh and an energy cutoff of 300 eV.
The calculated γα/β with the increase in atomic number is further depicted in Figure 3b–d. According to the CNT, the theoretical nucleation radius R* cannot be calculated by a negative γα/β, and the γα/β of all Al/Al3TM are less than 0 mJ·m−2, except for the (111) of Al/Al3Sc, Al/Al3Ti, Al/Al3(Y-Zr) and Al3Hf systems.
It can be seen from Figure 3b,c that the γα/β of Al/Al3TM (TM = (Sc-Zn, Y-Cd)) decreases from Sc and Y to Mn and Tc, and then increases to Zn and Cd, respectively, except for the (001) of Al/Al3(Fe-Co), the (111) of Al/Al3Pd and the (110) and (111) of Al/Al3Cd. These trends of γα/β for Al/Al3TM (TM = (Hf-Hg)) in the (110) and (111) systems present two Al/Al3Re and Al/Al3Pt compound troughs in Figure 3d, and they show the same change with the increase in atomic number. For the (001) system, the γα/β of Al/Al3TM (TM = (Hf-Hg) is larger than −250 mJ·m−2, and the Al/Al3TM with 3d64s2 has the lowest γα/β. Figure 3e shows the variation of strain energy ΔEcs of Al/Al3TM (TM = (Sc-Zn) with the increase in atomic number. It can be seen that the ΔEcs increases from 0.32~1.52 meV·atom−1 for Sc to 12.06 meV·atom−1 for Co on the (001) system, to 16.58 meV·atom−1 for Fe on the (110) system, and to 28.62 meV·atom−1 for Mn on the (111) system, respectively, and then they all decrease to 0.24 ~ 0.67 meV·atom−1 for Zn (except for Al/Al3Mn, of the order of −0.14 meV·atom−1). For the (110) and (111) systems of Al/Al3TM (TM = (Y-Cd, Hf-Hg)), as seen in Figure 3f,g, respectively, the largest values of ΔEcs for the (110) and (111) interface systems are all located at Al/Al3Re, being 26.49 and 33.54 meV·atom−1, respectively, while the (001) interface system of Al/Al3Tc has the lowest value of ΔEcs, being −14.84 meV·atom−1.
The trends of ΔGV as a function of temperature for all Al3TM compounds have been calculated according to Equations (10) and (11), and results are shown as Figure 3h. The results show that the ΔGV of all Al3TM change slightly in the temperature range of 0~1000 K, except that the ΔGV of Al3Sc, Al3Cu, Al3(Y-Zr), Al3Cd, Al3Hf and Al3Hg increase nonlinearly from −89.69, −1.44, −130.51, −93.86, −1.65, −72.35, and 0.65 meV·atom−1 to −24.38, 66.09, −88.46, −44.47, 71.60, −2.05 and ~88.51 meV·atom−1, respectively. Furthermore, the obtained ΔGV as a function of the atomic number of TM is shown in Figure 3i,j, and the calculated values of −66.46 and −61.54 meV·atom−1 for Al3Sc and Al3Ti, respectively, at 600 K agree well with the value of −61.14 meV·atom−1 at 350 °C (623 K) for Al3Sc and −66.15 meV·atom−1 at 300 (573 K) for Al3Ti calculated by Li et al. [15]. From Figure 3i, one can see that the ΔGV at 300 K increases from −80.96 meV·atom−1 for Sc, −120.46 meV·atom−1 for Y and −66.82 meV·atom−1 for Hf to 20.37 meV·atom−1 for Mn, 53.89 meV·atom−1 for Tc and 74.50 meV·atom−1 for Re, and then decreases slightly to −11.72 meV·atom−1 for Co, 9.62 meV·atom−1 for Rh and 4.89 meV·atom−1 for Ir, respectively. As a final step, they change slightly. At 600 K, the variation trends of ΔGV for 3–5d TMs are the same as those at 300 K.

3.3. Surface Energy

In the framework of Peierls theory, a lower surface energy of bulk materials in comparison to an unstable stacking fault will cause metals to crack from material failure [42,60]. Thus, it is necessary to analyze surface energy for all Al3TM particles and the Al matrix. The surface energy of Al is given by the following formula [61,62]:
E s u r = E A l s l a b N μ A l b u l k 2 A
where E A l s l a b and N are the total energy and the number of Al atoms in the slab model, respectively. μ A l b u l k represents the chemical potential of a single atom in bulk Al.
For stoichiometric surfaces (111) of the Al3TM slab, the calculated formula is given as follows:
3 μ A l s l a b + μ T M s l a b = μ A l 3 T M b u l k
E s u r = E A l 3 T M s l a b N μ A l 3 T M b u l k 2 A
where μ A l s l a b , μ T M s l a b and μ A l 3 T M b u l k are the chemical potential of Al, AlTM-terminated and Al3TM bulk, respectively. N and A are the number of Al3TM cells and the surface area, respectively.
To further discuss the non-stoichiometric (001) and (110) surfaces of the Al3TM (3NTM ≠ NAl), we used the following the equation [24,63]:
E s u r = E A l 3 T M s l a b N μ A l 3 T M b u l k + n μ A l s l a b 2 A
where n is the number of the rest (n < 0) and missing (n > 0) Al atoms.
To obtain μ A l s l a b in the systems, we first need to avoid Al and Sc bulk phases. Therefore, μ A l s l a b and μ S c s l a b are limited, as follows:
μ A l s l a b μ A l b u l k < 0
μ T M s l a b μ T M b u l k < 0
Further, the thermodynamic stability of AlTM compounds should meet the equation given by:
3 μ A l b u l k + μ T M b u l k + H f = μ A l 3 T M b u l k
Combined with Equations (12) and (15)–(17), two limit values of μ A l s l a b are respectively given by:
μ A l a = μ A l s l a b = μ A l b u l k
μ A l b = μ A l s l a b = μ A l b u l k + 1 3 H f
The calculated surface energies of AlTM from different Al chemical potentials, μ A l a and μ A l b , are summarized in Table 3 as references. To solve the dependence of surface energy on Al chemical potential, the average surface energy of non-stoichiometric surfaces is obtained by two identical index surfaces of different termination [24,63]:
E s u r a v e = 1 4 A E s l a b A l + E s l a b A l T M N s l a b A l + N s l a b A l T M × μ A l 3 T M b u l k
where E s l a b A l , N s l a b A l and E s l a b A l T M , N s l a b A l T M are the relaxed energy and total number of TM atoms in Al and AlTM-terminated surfaces, respectively.
Figure 4a shows the slab model for calculating Esur, and the detail calculated results of E s u r a v e of Al3TM (RE = Sc-Zn, Y-Cd and Hf-Hg) are depicted in Table 4. It can be seen that for pure Al, Al3(Sc-V) and Al3(Y-Nb), the calculated values of this work are in good agreement with references [24,64,65]. The E s u r a v e of low index surfaces of cubic Al follows in the sequence of (110) > (100) > (111), which follows the general law of surface energies for face-centered cubic metals [66]. Figure 4c–e illustrates the change in E s u r a v e with the atomic number of TM elements, and it can be found that the variation tendency of E s u r a v e of Al3TM intermetallic compounds presents similar characteristics for different cycles. For example, the E s u r a v e of Al3TM for 3d elements firstly decreases from Sc to Mn, and then increases to Fe, and then decreases to Ni, and finally changes slightly in the (001) surface. The variation ranges of E s u r a v e for the (001), (110) and (111) surfaces of Al3TM are 0.25~1.44, 0.26~1.45 and −0.32~1.18 J·m−2, respectively, and the (111) surface has the lowest surface energy for all elements. As can be seen from Figure 4e,f, we have calculated the values of ELF by using the Equation (7) on the (111) plane of Al3Sc and Al3Mo. It can be seen that the (111) plane of Al3Sc has more strongly localized electron areas than the (111) plane of Al3Mo, indicating that strongly localized electrons make the surface energy lower. Clearly, if the E s u r a v e of Al3TM is larger than that of Al, they would increase the toughness of Al alloys such as Al3Sc. However, the complete toughness is not only determined by the above results but also by assessing the information of generalized stacking fault energy (GSFE) for particles in the Al matrix. The work needed for this is underway and will be published elsewhere.

4. Conclusions

In this work, we have calculated the diffusion rates of TM elements in Al and the key parameters of nucleation and growth of second phase particles Al3TM (TM = Sc-Zn, Y-Cd and Hf-Hg) using the first principles combing with quasi-harmonic approximation in the theoretical framework of density functional theory. Firstly, as can be seen from the discussed results, the trends of the Q, chemical formation and surface energies with the change in atomic number in the same cycle are similar. The reason may be related to the valence electron configures (VECs) of TM elements because TM in the same cycle has the same VECs. Meanwhile, the calculation interface and strain energy of Al3TM/Al composed of the same cycle of TM elements show a lack of similarity. The reason for this may be that the interface and strain energy are co-determined by the matrix and second phases. Here, the main conclusions are as follows:
  • In the vacancy–substitution model, the diffusion activation energy Q first increases, and then decreases with the increase in atomic number (Sc-Zn, Y-Ag and Hf-Au) in the Al matrix, except for Cr; the TM elements in the fourth cycle generally have lower Qs.
  • Mn-Co, Tc-Rh and Re-Ir elements have larger activation energies Qs in the Al matrix, while Cu-Zn, Ag and Au have lower activation energies Qs; even Cd and Hg elements have negative activation energies. In the undoped-Al system, the self-diffusion activation energy Q0 is lower compared to all Qs in the doped system, except for the Qs of Cu, Zn, Y and Ag.
  • The diffusion rate Ds of all impurity atoms increases logarithmically with the increase in temperature. With the increase in atomic number, the diffusion rate Ds first decreases linearly from Sc, Y and Hf to Mn, Ru and Ir, and then increases to Zn, Ag and Au for 3–5d TM elements, respectively.
  • With the increase in atomic number, the interface energy γα/β of Al/Al3TM (TM = (Sc-Zn, Y-Cd)) decreases from Sc and Y to Mn and Tc, and then increases to Zn and Cd, respectively, except for (001) in Al/Al3(Fe-Co), (111) in Al/Al3Pd and (110) and (111) in Al/Al3Cd. Meanwhile, the strain energy ΔEcs increases from Sc to Co in the (001) system, to Fe in the (110) system, and to Mn in the (111) system, respectively, and then they all decreases to Zn, except for Al/Al3Mn. The largest values of ΔEcs for (110) and (111) interface systems are all located at Al/Al3Re, while the (001) interface system of Al/Al3Tc has the lowest value.
  • The variation in chemical formation energy ΔGV of all Al3TM changes slightly in the temperature range of 0~1000 K, except that the ΔGV of Al3Sc, Al3Cu, Al3(Y-Zr), Al3Cd, Al3Hf and Al3Hg increase nonlinearly. With the increase in atomic number at 300 K, the ΔGV increases from Sc, Y and Hf to Mn, Tc and Re at first, and then decreases to Co, Rh and Ir, respectively, and finally, it slightly changes. The variation trends of the ΔGV for 3–5d TMs are the same as those at 300 K.
  • With the increase in atomic number, the trend of E s u r a v e of Al3TM intermetallic compounds presents a similar change in different cycles and the (111) surface always has the lowest surface energy in all surfaces of Al3TM particles.

Author Contributions

Methodology, T.F.; Investigation, Z.R. and T.F.; Writing—original draft, Z.R.; Writing—review & editing, T.F.; Visualization, T.H., Z.R., K.W., K.H. and Y.W.; Project administration, T.H., K.W. and Y.W. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by the Foshan Technology Project (1920001000409), the Scientific Research Project of Hunan Institute of Technology (HQ21016, 21A0564, HP21047, 21B0796), the Natural Science Foundation of China (52171115). The APC was funded by Foshan Technology Project (1920001000409).

Data Availability Statement

In this article, the key parameters of growth and nucleation of Al3TM particles based on the corresponding models (see Figure 2 and Figure 3) have been discussed in our work. Figure 4 describe calculated surface energy of Al3TM and Al. Table 1, Table 2, Table 3 and Table 4 present important data as a reference. All data can be found in the manuscript.

Conflicts of Interest

The authors declare that they have no known competing financial interest or personal relationships that have influenced the work reported in this paper.

References

  1. Schlapbach, L.; Züttel, A. Hydrogen-storage materials for mobile applications. Nature 2001, 414, 353–358. [Google Scholar] [CrossRef] [PubMed]
  2. Wen, K.; Xiong, B.-Q.; Fan, Y.-Q.; Zhang, Y.-A.; Li, Z.-H.; Li, X.-W.; Wang, F.; Liu, H.-W. Transformation and dissolution of second phases during solution treatment of an Al-Zn-Mg-Cu alloy containing high zinc. Rare Met. 2018, 37, 376–380. [Google Scholar] [CrossRef]
  3. 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]
  4. Drits, M.E.; Kadaner, E.S.; Turkina, N.I.; Fedotov, S.G. The Mechanical Properties of Aluminum-Lithium Alloy. Transl. Splav. Tsvetn. Met. 1972, 14. [Google Scholar]
  5. Yang, Y.; Licavoli, J.J.; Hackney, S.A.; Sanders, P.G. Coarsening behavior of precipitate Al3(Sc, Zr) in supersaturated Al-Sc-Zr alloy via melt spinning and extrusion. J. Mater. Sci. 2021, 56, 11114–11136. [Google Scholar] [CrossRef]
  6. Yan, K.; Chen, Z.; Lu, W.J.; Zhao, Y.; Le, W.; Naseem, S. Nucleation and growth of Al3Sc precipitates during isothermal aging of Al-0.55wt% Sc alloy. Mater. Charact. 2021, 179, 111331. [Google Scholar] [CrossRef]
  7. Clemens, H.; Kestler, H. Processing and Applications of Intermetallic -TiAl-Based Alloys. Adv. Eng. Mater. 2010, 2, 551–570. [Google Scholar] [CrossRef]
  8. Ug, Ş.; Arıkan, N.; Soyalp, F.; Ug, G. Phonon and elastic properties of AlSc and MgSc from first-principles calculations. Comput. Mater. Sci. 2010, 48, 866–870. [Google Scholar]
  9. RMichi, A.; Toinin, J.P.; Farkoosh, A.R.; Seidman, D.N.; Dunand, D.C. Effects of Zn and Cr additions on precipitation and creep behavior of a dilute Al–Zr–Er–Si alloy. Acta Mater. 2019, 181, 249–261. [Google Scholar]
  10. Zedalis, M.S.; Fine, M.E. Precipitation and ostwald ripening in dilute AI Base-Zr-V alloys. Metall. Trans. A 1986, 17, 2187–2198. [Google Scholar] [CrossRef]
  11. Parameswaran, V.R.; Weertman, J.R.; Fine, M.E. Coarsening behavior of L12 phase in an Al-Zr-Ti alloy. Scr. Metall. 1989, 23, 147–150. [Google Scholar] [CrossRef]
  12. Chen, Z.; Zhang, P.; Chen, D.; Wu, Y.; Wang, M.; Ma, N.; Wang, H. First-principles investigation of thermodynamic, elastic and electronic properties of Al3V and Al3Nb intermetallics under pressures. J. Appl. Phys. 2015, 117, 085904. [Google Scholar] [CrossRef]
  13. Li, R.-Y.; Duan, Y.-H. Electronic structures and thermodynamic properties of HfAl3 in L12, D022 and D023 structures. Trans. Nonferrous Met. Soc. China 2016, 26, 2404–2412. [Google Scholar] [CrossRef]
  14. Czerwinski, F. Thermal Stability of Aluminum Alloys. Materials 2020, 13, 3441. [Google Scholar] [CrossRef] [PubMed]
  15. 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]
  16. Zhang, X.; Huang, Y.; Liu, Y.; Ren, X. A comprehensive DFT study on the thermodynamic and mechanical properties of L12-Al3Ti/Al interface. Vacuum 2021, 183, 109858. [Google Scholar] [CrossRef]
  17. Wang, Y.; Wang, J.; Zhang, C.; Huang, H. Mechanical properties of defective L12-Al3X (X= Sc, Lu) phase: A first-principles study. J. Rare Earths 2021, 39, 217–224. [Google Scholar] [CrossRef]
  18. Dorin, T.; Babaniaris, S.; Jiang, L.; Cassel, A.; Robson, J.D. Stability and stoichiometry of L12 Al3(Sc,Zr) dispersoids in Al-(Si)-Sc-Zr alloys. Acta Mater. 2021, 216, 117117. [Google Scholar] [CrossRef]
  19. Liu, T.; Ma, T.; Li, Y.; Ren, Y.; Liu, W. Stable mechanical and thermodynamic properties of Al-RE intermetallics: A First-principles study. J. Rare Earths 2022, 40, 345–352. [Google Scholar] [CrossRef]
  20. Nakai, M.; Eto, T. New aspect of development of high strength aluminum alloys for aerospace applications. Mater. Sci. Eng. A 2000, 285, 62–68. [Google Scholar] [CrossRef]
  21. Saha, S.; Todorova, T.Z.; Zwanziger, J.W. Temperature dependent lattice misfit and coherency of Al3X (X=Sc, Zr, Ti and Nb) particles in an Al matrix. Acta Mater. 2015, 89, 109–115. [Google Scholar] [CrossRef]
  22. 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]
  23. 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]
  24. Sun, S.P.; Li, X.P.; Wang, H.J.; Jiang, H.F.; Lei, W.N.; Jiang, Y.; Yi, D.Q. First-principles investigations on the electronic properties and stabilities of low-index surfaces of L12–Al3Sc intermetallic. Appl. Surf. Sci. 2014, 288, 609–618. [Google Scholar] [CrossRef]
  25. Hood, G.M. The diffusion of iron in aluminium. Philos. Mag. 1970, 21, 305–328. [Google Scholar] [CrossRef]
  26. Mantl, S.; Petry, W.; Schroeder, K.; Vogl, G. Diffusion of iron in aluminum studied by Mössbauer spectroscopy. Phys. Rev. B 1983, 27, 5313–5331. [Google Scholar] [CrossRef]
  27. 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]
  28. Mandal, P.K.; Kumar, R.J.F.; Varkey, J.M. Effect of artificial ageing treatment and precipitation on mechanical properties and fracture mechanism of friction stir processed MgZn2 and Al3Sc phases in aluminium alloy. Mater. Today Proc. 2021, 46, 4982–4987. [Google Scholar] [CrossRef]
  29. Alexander, W.B.; Slifkin, L.M. Diffusion of Solutes in Aluminum and Dilute Aluminum Alloys. Phys. Rev. B 1970, 1, 3274–3282. [Google Scholar] [CrossRef]
  30. Zhao, X.; Chen, H.; Wilson, N.; Liu, Q.; Nie, J.-F. Direct observation and impact of co-segregated atoms in magnesium having multiple alloying elements. Nat. Commun. 2019, 10, 3243. [Google Scholar] [CrossRef] [Green Version]
  31. Maruhn, J.A.; Reinhard, P.G.; Suraud, E. Density Functional Theory; Springer: Berlin/Heidelberg, Germany, 2010. [Google Scholar]
  32. Finnis, M.W. The theory of metal-ceramic interfaces. J. Phys. Condens. Matter 1996, 8, 5811–5836. [Google Scholar] [CrossRef]
  33. Zhou, W.F.; Ren, X.D.; Ren, Y.P.; Yuan, S.Q.; Ren, N.F.; Yang, X.Q.; Adu-Gyamfi, S. Initial dislocation density effect on strain hardening in FCC aluminium alloy under laser shock peening. Philos. Mag. 2017, 97, 917–929. [Google Scholar] [CrossRef]
  34. Zhao, S.J.; Stocks, G.M.; Zhang, Y.W. Stacking fault energies of face-centered cubic concentrated solid solution alloys. Acta Mater. 2017, 134, 334–345. [Google Scholar] [CrossRef]
  35. Mardirossian, N.; Head-Gordon, M. Thirty years of density functional theory in computational chemistry: An overview and extensive assessment of 200 density functionals. Mol. Phys. 2017, 115, 2315–2372. [Google Scholar] [CrossRef] [Green Version]
  36. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758–1775. [Google Scholar] [CrossRef]
  37. Hafner, J. Ab-initio simulations of materials using VASP: Density-functional theory and beyond. J. Comput. Chem. 2008, 29, 2044–2078. [Google Scholar] [CrossRef]
  38. TFischer, H.; Almlof, J. General methods for geometry and wave function optimization. J. Phys. Chem. 1992, 96, 9768–9774. [Google Scholar] [CrossRef]
  39. Liu, Y.; Wen, J.C.; Zhang, X.Y.; Huang, Y.C. A comparative study on heterogeneous nucleation and mechanical properties of the fcc-Al/L12-Al3M (M = Sc, Ti, V, Y, Zr, Nb) interface from first-principles calculations. Phys. Chem. Chem. Phys. 2021, 23, 4718–4727. [Google Scholar] [CrossRef]
  40. 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] [Green Version]
  41. Lifshitz, I.M.; Slyozov, V.V. The kinetics of precipitation from supersaturated solid solutions. J. Phys. Chem. Solids 1961, 19, 35–50. [Google Scholar] [CrossRef]
  42. Wagner, C. Zeitschrift für Elektrochemie. Berichte der Bunsengesellschaft für physikalische Chemie. 1961, 65, 7–8. [Google Scholar]
  43. Mantina, M.; Wang, Y.; Arroyave, R. First-Principles Calculation of Self-Diffusion Coefficients. Phys. Rev. Lett. 2008, 100, 215901. [Google Scholar] [CrossRef] [Green Version]
  44. Mantina, M.; Wang, Y.; Chen, L.Q.; Liu, Z.K.; Wolverton, C. First principles impurity diffusion coefficients. Acta Mater. 2009, 57, 4102–4108. [Google Scholar] [CrossRef]
  45. Evangelakis, G.; Papanicolaou, N. Adatom self-diffusion processes on (001) copper surface by molecular dynamics. Surf. Sci. 1996, 347, 376–386. [Google Scholar] [CrossRef]
  46. Gomer, R. Diffusion of adsorbates on metal surfaces. Rep. Prog. Phys. 1990, 53, 917. [Google Scholar] [CrossRef] [Green Version]
  47. Bennett, C.H. Exact defect calculations in model substances. Diffus. Solids Recent Dev. 1975, 1, 73–113. [Google Scholar]
  48. Vineyard, G.H. Frequency factors and isotope effects in solid state rate processes. J. Phys. Chem. Solids 1957, 3, 121–127. [Google Scholar] [CrossRef]
  49. Yang, B.; Wang, L.-G.; Yi, Y.; Wang, E.-Z.; Peng, L.-X. First-principles calculations of the diffusion behaviors of C, N and O atoms in V metal. Acta Phys. Sin. 2015, 64, 026602. [Google Scholar] [CrossRef]
  50. Chen, L.; Li, Y.; Xiao, B.; Gao, Y.; Zhao, S. Chemical bonding, thermodynamic stability and mechanical strength of Ni3Ti/α-Al2O3 interfaces by first-principles study. Scr. Mater. 2021, 190, 57–62. [Google Scholar] [CrossRef]
  51. Wert, C. Diffusion coefficient of C in α-iron. Phys. Rev. 1950, 79, 601. [Google Scholar] [CrossRef]
  52. Wert, C.; Zener, C. Interstitial atomic diffusion coefficients. Phys. Rev. 1949, 76, 1169. [Google Scholar] [CrossRef]
  53. Liu, P.; Wang, S.; Li, D.; Li, Y.; Chen, X.-Q. Fast and Huge Anisotropic Diffusion of Cu (Ag) and Its Resistance on the Sn Self-diffusivity in Solid β–Sn. J. Mater. Sci. Technol. 2016, 32, 121–128. [Google Scholar] [CrossRef]
  54. Ehrhart, P. Atomic Defects in Metals; Springer: Berlin/Heidelberg, Germany, 1991. [Google Scholar]
  55. Fujikawa, S.I. Impurity Diffusion of Scandium in Aluminium. Defect Diffus. Forum 1997, 143, 115–120. [Google Scholar] [CrossRef]
  56. Volin, T.E.; Balluffi, R.W. Annealing kinetics of voids and the Self-diffusion coefficient in aluminum. Phys. Status Solidi 2010, 25, 163–173. [Google Scholar] [CrossRef]
  57. Murphy, J.B. Interdiffusion in dilute aluminium-copper solid solutions. Acta Metall. 1961, 9, 563–569. [Google Scholar] [CrossRef]
  58. Feng, Y.; Liu, M.; Shi, Y.; Ma, H.; Li, D.; Li, Y.; Lu, L.; Chen, X.Q. High-throughput modeling of atomic diffusion migration energy barrier of fcc metals. Prog. Nat. Sci. Mater. Int. 2019, 29, 341–348. [Google Scholar] [CrossRef]
  59. Pareige, C.; Soisson, F.; Martin, G.; Blavette, D. Ordering and phase separation in Ni–Cr–Al: Monte Carlo simulations vs. three-dimensional atom probe. Acta Mater. 1999, 47, 1889–1899. [Google Scholar] [CrossRef]
  60. Marder, M. Correlations and Ostwald ripening. Phys. Rev. A 1987, 36, 858. [Google Scholar] [CrossRef]
  61. Hirth, J.P.; Lothe, J. Theory of Dislocations, 2nd ed.; Wiley: New York, NY, USA, 1982. [Google Scholar]
  62. Hutchinson, J.W. Singular behavior at the end of a tensile crack in a hardening material. J. Mech. Phys. Solids 1968, 16, 13–31. [Google Scholar] [CrossRef]
  63. Rice, J.R.; Rosengren, G.F. Plane strain deformation near a crack tip in a power-law hardening material. J. Mech. Phys. Solids 1968, 16, 1–12. [Google Scholar] [CrossRef]
  64. Zhang, X.; Ren, X.; Li, H.; Zhao, Y.; Huang, Y.; Liu, Y.; Xiao, Z. Interfacial properties and fracture behavior of the L12-Al3Sc || Al interface: Insights from a first-principles study. Appl. Surf. Sci. 2020, 515, 146017. [Google Scholar] [CrossRef]
  65. Wang, Y.X.; Arai, M.; Sasaki, T.; Wang, C.L. First-principles study of the (001) surface of cubic CaTiO3. Phys. Rev. B 2006, 73, 035411. [Google Scholar] [CrossRef]
  66. Vitos, L.; Ruban, A.V.; Skriver, H.L.; Kollár, J. The surface energy of metals. Surf. Sci. 1998, 411, 186–202. [Google Scholar] [CrossRef]
Figure 1. The total energy of Al3Sc as a function of k-point sampling grids.
Figure 1. The total energy of Al3Sc as a function of k-point sampling grids.
Crystals 13 01032 g001
Figure 3. (a) The interface models. (bd) The calculated interface energy γα/β and (eg) strain energy ΔEcs with the change in atomic number. (h) The chemical formation energy ΔGV with the change in temperature. (i,j) The chemical formation energy ΔGV as a function of atomic number under different constant temperature conditions. (a) represent the calculated result of Li et al. [23].
Figure 3. (a) The interface models. (bd) The calculated interface energy γα/β and (eg) strain energy ΔEcs with the change in atomic number. (h) The chemical formation energy ΔGV with the change in temperature. (i,j) The chemical formation energy ΔGV as a function of atomic number under different constant temperature conditions. (a) represent the calculated result of Li et al. [23].
Crystals 13 01032 g003
Figure 4. (a) The slab model. (bd) The calculated average surface energy E s u r a v e with the change in atomic number. (e,f) The ELF on the (111) plane of Al3Sc and Al3Mo systems, respectively.
Figure 4. (a) The slab model. (bd) The calculated average surface energy E s u r a v e with the change in atomic number. (e,f) The ELF on the (111) plane of Al3Sc and Al3Mo systems, respectively.
Crystals 13 01032 g004
Table 2. The calculated interface energy γα/β (mJ·m−2) and strain energy ΔEcs (meV·atom−1) in Al/Al3TM interface systems. (Note: A * symbol represents the calculated result from the vacuum slab model).
Table 2. The calculated interface energy γα/β (mJ·m−2) and strain energy ΔEcs (meV·atom−1) in Al/Al3TM interface systems. (Note: A * symbol represents the calculated result from the vacuum slab model).
Dir.(001)(110)(111)
Systemsγα/βΔECSγα/βΔECSγα/βΔECS
Al/Al3Sc108.65
108.00 [15]
165.00 [23]
176.00 [23]
0.32
0.60 [23]
-
-
194.41
159.00 [15]
178.00 [15]
193.00 [23]
1.50204.81
191.00 [23]
189.00 [15]
203.00 [15]
1.52
Al/Al3Ti−38.48
61.85 *
52.00 [23]
0.20
0.30 [15]
−38.90
61.00
1.1266.67
79.00 [15]
1.36
Al/Al3V−147.831.77−203.775.10−75.767.56
Al/Al3Cr−270.041.52−379.959.92−167.4323.16
Al/Al3Mn−468.86−0.14−429.6513.80−225.6628.62
Al/Al3Fe−291.2210.64−283.3716.58−113.5423.69
Al/Al3Co−200.0412.06−205.3114.01−49.4922.43
Al/Al3Ni−195.595.83−176.896.20−109.5015.46
Al/Al3Cu−143.590.59−108.272.40−48.538.43
Al/Al3Zn−53.810.64−88.330.67−33.480.24
Al/Al3Y93.375.13159.609.72181.2914.30
Al/Al3Zr20.150.651.392.4386.322.48
Al/Al3Nb−143.961.48−160.592.07−109.570.56
Al/Al3Mo−309.651.87−319.591.20−201.0418.33
Al/Al3Tc−699.48−14.84−516.467.24−201.1020.79
Al/Al3Ru−173.703.92−228.408.52−82.7112.26
Al/Al3Rh−138.072.92−197.666.71−42.729.98
Al/Al3Pd−132.751.53−153.881.94−60.521.79
Al/Al3Ag−142.470.86−46.851.31−5.670.40
Al/Al3Cd−75.680.51−261.594.33−55.699.33
Al/Al3Hf−37.531.14−25.311.7969.411.59
Al/Al3Ta−169.680.50−198.591.36−124.381.26
Al/Al3W−232.180.29−467.974.35−276.164.24
Al/Al3Re−146.354.59−1242.0026.49−396.5733.54
Al/Al3Os−243.804.34−328.808.72−174.0613.36
Al/Al3Ir−87.863.71−173.406.26−3.4410.60
Al/Al3Pt−190.370.38−734.312.90−246.3917.81
Al/Al3Au−118.520.58−80.771.02−33.300.67
Al/Al3Hg−93.950.94−341.483.64−251.4024.04
Table 3. The calculated surface energy Esur (J·m−2) of the (001) and (110) surfaces from different Al chemical potential μ A l a and μ A l b in Al3TM.
Table 3. The calculated surface energy Esur (J·m−2) of the (001) and (110) surfaces from different Al chemical potential μ A l a and μ A l b in Al3TM.
Systems(001)(110)
Al-Ter.AlTM-Ter.Al-Ter.AlTM-Ter.
μ A l a μ A l b μ A l a μ A l b μ A l a μ A l b μ A l a μ A l b
Al3Sc1.10 1.69 1.42 0.84 1.19 1.61 1.63 1.22
Al3Ti1.04 1.53 1.70 1.21 0.99 1.34 1.84 1.49
Al3V1.03 1.24 1.61 1.40 0.86 1.01 1.77 1.63
Al3Cr0.97 0.92 1.54 1.59 0.63 0.60 1.50 1.53
Al3Mn0.93 0.99 1.55 1.49 0.42 0.47 1.19 1.14
Al3Fe1.06 1.23 1.57 1.39 0.78 0.90 1.51 1.39
Al3Co0.97 1.29 1.32 0.99 0.77 0.99 1.32 1.09
Al3Ni0.78 1.10 1.03 0.71 0.73 0.96 0.90 0.68
Al3Cu0.83 0.89 0.99 0.92 0.87 0.92 0.85 0.80
Al3Zn0.84 0.82 0.78 0.80 0.93 0.91 0.75 0.77
Al3Y1.13 1.63 0.96 0.46 1.12 1.47 1.34 0.98
Al3Zr0.88 1.46 1.36 0.77 0.88 1.29 1.59 1.18
Al3Nb0.81 1.16 1.32 0.97 0.70 0.95 1.61 1.37
Al3Mo0.36 0.51 0.88 0.73 0.26 0.36 1.17 1.08
Al3Tc1.02 1.29 1.62 1.35 0.00 0.18 0.77 0.59
Al3Ru1.17 1.67 1.62 1.11 0.80 1.15 1.53 1.18
Al3Rh0.85 1.54 1.01 0.31 0.66 1.15 1.04 0.55
Al3Pd0.58 1.05 0.67 0.20 0.44 0.80 0.47 0.12
Al3Ag0.68 0.63 0.53 0.58 0.72 0.68 0.50 0.53
Al3Cd0.81 0.63 0.23 0.41 0.72 0.60 0.40 0.53
Al3Hf0.96 1.45 1.63 1.14 0.94 1.29 1.78 1.43
Al3Ta0.91 1.10 1.66 1.46 0.75 0.89 1.82 1.68
Al3W0.64 0.54 1.45 1.54 0.57 0.50 1.50 1.56
Al3Re0.90 0.95 1.76 1.70 0.00 0.04 0.93 0.89
Al3Os1.08 1.40 1.81 1.49 0.59 0.82 1.52 1.29
Al3Ir1.07 1.75 1.38 0.69 0.77 1.26 1.39 0.91
Al3Pt0.66 1.30 0.71 0.07 0.49 0.97 0.55 0.08
Al3Au0.55 0.70 0.35 0.20 0.46 0.57 0.31 0.20
Al3Hg0.67 0.45 −0.15 0.07 0.35 0.20 0.19 0.34
Table 4. The calculated surface energy Esur (J·m−2) in Al or Al3TM.
Table 4. The calculated surface energy Esur (J·m−2) in Al or Al3TM.
Systems(001) (110) (111)
Al0.79; 0.93 [23]0.87; 0.98 [23]0.68; 0.73 [39] 0.81 [23]
Al3Sc1.26; 1.32 [24]1.41; 1.45 [24]1.18; 1.22 [24]; 1.17 [39]
Al3Ti1.37 1.42 0.92; 0.93 [39]
Al3V1.32 1.32 0.72; 0.65 [39]
Al3Cr1.25 1.07 0.33
Al3Mn1.24 0.81 0.33
Al3Fe1.31 1.15 0.56
Al3Co1.14 1.04 0.67
Al3Ni0.91 0.82 0.57
Al3Cu0.91 0.86 0.69
Al3Zn0.81 0.84 0.73
Al3Y1.051.231.06; 1.11 [39]
Al3Zr1.12 1.23 0.80; 0.94 [39]
Al3Nb1.06 1.16 0.54; 0.59 [39]
Al3Mo0.62 0.73 −0.32
Al3Tc1.32 0.39 −0.22
Al3Ru1.39 1.17 0.44
Al3Rh0.93 0.85 0.46
Al3Pd0.62 0.46 0.21
Al3Ag0.61 0.61 0.46
Al3Cd0.52 0.56 0.47
Al3Hf1.30 1.36 0.87
Al3Ta1.28 1.29 0.66
Al3W1.04 1.05 0.13
Al3Re1.32 0.46 −0.32
Al3Os1.45 1.06 0.41
Al3Ir1.22 1.08 0.73
Al3Pt0.69 0.52 0.30
Al3Au0.45 0.38 0.35
Al3Hg0.26 0.27 0.27
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

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. https://doi.org/10.3390/cryst13071032

AMA Style

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(7):1032. https://doi.org/10.3390/cryst13071032

Chicago/Turabian Style

Hu, Te, Zixiong Ruan, Touwen Fan, Kai Wang, Kuanfang He, and Yuanzhi Wu. 2023. "First-Principles Investigation of the Diffusion of TM and the Nucleation and Growth of L12 Al3TM Particles in Al Alloys" Crystals 13, no. 7: 1032. https://doi.org/10.3390/cryst13071032

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