Next Article in Journal
Degradation Analysis of Thin Mg-xAg Wires Using X-ray Near-Field Holotomography
Previous Article in Journal
Advanced High-Strength Steels by Quenching and Partitioning
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Thermal Properties of L12 Phases in Aluminum Enhanced by Alloying Elements

1
Hunan Province Key Laboratory of Materials Surface & Interface Science and Technology, Central South University of Forestry and Technology, Changsha 410004, China
2
College of Science, Central South University of Forestry and Technology, Changsha 410004, China
3
Center for High-Resolution Electron Microscopy, College of Materials Science and Engineering, Hunan University, Changsha 410082, China
4
Guangdong Key Laboratory for Hydrogen Energy Technologies, School of Materials Science and Hydrogen Energy, Foshan University, Foshan 528000, China
*
Authors to whom correspondence should be addressed.
Metals 2021, 11(9), 1420; https://doi.org/10.3390/met11091420
Submission received: 4 July 2021 / Revised: 5 September 2021 / Accepted: 6 September 2021 / Published: 8 September 2021

Abstract

:
The L12 type trialuminide compounds Al3M possess outstanding mechanical properties, which enable them to be ideal for dispersed strengthening phases for the high-strength thermally stable Al based alloys. Ab-initio calculations based on the density functional theory (DFT) were performed to study the structural, electronic, thermal, and thermodynamic properties of L12-Al3M (M = Er, Hf, Lu, Sc, Ti, Tm, Yb, Li, Mg, Zr) structures in Al alloys. The total energy calculations showed that the L12 structures are quite stable. On the basis of the thermodynamic calculation, we found that the Yb, Lu, Er, and Tm atoms with a larger atomic radii than Al promoted the thermal stability of the Al alloys, and the thermal stability rank has been constructed as: Al3Yb > Al3Lu > Al3Er > Al3Tm > Al, which shows an apparent positive correlation between the atomic size and thermal stability. The chemical bond offers a firm basis upon which to forge links not only within chemistry but also with the macroscopic properties of materials. A careful analysis of the charge density indicated that Yb, Lu, Er, and Tm atoms covalently bonded to Al, providing a strong intrinsic basis for the thermal stability of the respective structures, suggesting that the addition of big atoms (Yb, Lu, Er, and Tm) are beneficial for the thermal stability of Al alloys.

1. Introduction

The considerable attention that aluminum alloys have received originates from their excellent physical and chemical properties, such as low density, high specific strength, easy processing and molding, excellent electrical conductivity, thermal conductivity, corrosion resistance, etc. Thus, Al alloys are nowadays widely used in the aviation, aerospace, automotive, machinery manufacturing, shipping, and chemical industries [1,2,3]. However, scientific development and advanced technological projects, e.g., high-speed trains, and large aircraft and spacecraft, put higher requirements in front of the overall performance of aluminum alloys.
The structure of the L12 type compounds Al3M similarity with the Al matrix leads to a coherent interface, thereby minimizing the surface energy and maximizing the strengthening effect [4]. Recently, the L12-Al3M structure phase in Al alloy has received great attention since it was reported to improve the strength, toughness, wear resistance, corrosion resistance, fatigue resistance, and high-temperature resistance properties significantly [5,6]. Particularly, this structure was widely used as a high-temperature structural material because of its good thermal stability [7,8,9,10,11]. Al3Sc is the most extensively studied intermetallic compound because of its thermodynamically stable L12 structure, large volume fraction, and moderate coarsening rate [12,13,14]. However, the high cost of Sc limits its industrial applications. The alloying of RE elements (RE = Er, Tm, Yb, Lu) with Al yields thermodynamically stable L12-phase Al3RE precipitates, and RE elements are less expensive than Sc. Moreover, the substitution of Sc by RE elements also improved the high-temperature creep and hardness properties of Al3Sc [15,16]. Thus, Al3RE alloys are considered to be an important diffusion reinforcing phase for high-temperature applications, and the thermodynamic properties are investigated for understanding and predicting the response, thermal stability, and phase transformation of materials at high temperatures [4,9,15].
Investigations [4] have shown that many elements are potential additives to form the Al3M phases in Al alloys during aging heat treatment process, such as the four heaviest rare earth elements RE (RE = Er, Tm, Yb, Lu) [17], the Group IVA transition metal elements Ti, Zr and Hf, and other common elements in Al alloys. Previous studies of the L12-Al3M structure have focused on structural properties, mechanics, and thermodynamic properties (i.e., entropy and enthalpy formation). Zhang et al. [18] performed first-principles calculations on the thermodynamic properties of Al3Tm and Al3Lu under high pressure and found that Al3Tm exhibited better thermal stability than Al3Lu in the temperature range of 0–1000 K. Li et al. [19] calculated some of the thermal quantities of Al3Mg and Al3Sc alloys and confirmed that the quantities exhibited a linear high-temperature behavior at high pressure, eventually converging to an almost constant value at both high pressure and high temperature. Pan et al. [20] investigated the thermal properties and thermoelasticity of L12-Al3RE (RE = Er, Tm, Yb, Lu). It is found that the thermal properties of Al3RE are nearly isotropic in the considered temperature range, and the Al3RE phases can maintain a definite strength at a high temperature.
Although the L12-Al3M structure in Al alloys has attracted great attention and been widely investigated, there are fewer systematic experimental and theoretical studies on the thermodynamic properties of L12-Al3M structures, and the general rule of judging the thermodynamic properties of different L12-Al3M structures has not been performed up to now. In fact, structural information, thermodynamic properties, and bonding analysis are difficult to measure experimentally. With the development of computer science and technology, first-principle calculations have started to play an increasingly important role in predicting and analyzing materials’ physical and chemical properties.
Encouraged by the special interest of L12-Al3M structures in potential high strength Al alloys, eleven elements ( Er, Tm, Yb, Lu, Sc, Zr, Ti, Hf, Li, Mg, and Si) have been added to the Al matrix to analyze the thermodynamic properties using first-principles calculations systematically. Important physical quantities related to thermal properties, such as enthalpy of formation, Gibbs free energy, thermodynamic entropy, constant-pressure heat capacity, coefficient of thermal expansion, and Grüneisen parameters, are computed to reveal the thermodynamic properties of L12-Al3M structures. The purpose of the present paper is to show the structural features and the thermodynamic properties of individual L12-Al3M structures, and to try to propose an effective rule judging whether the addictive element is helpful to improve the therodynamic properties of Al alloys. Finally, we found that the addition of RE elements (Yb, Lu, Er, and Tm) significantly enhances the Al alloy thermal stability. The Al alloy thermal stability exhibits a significant positive correlation with the atomic radii of RE elements, which would positively affect the design and application of Al alloys in technology and industry.
Due to the limitation of small solid solubility and the diffusion coefficient of RE elements in the Al matrix [14,21,22], there are still some difficulties in experimental verification and testing. We believe that our results can provide strong theoretical support for subsequent experiments. We hope to overcome the present knowledge and manufacturing barriers and develop aluminum alloys with superior properties that remain stable under high temperatures.

2. Theoretical Methods

The first-principle calculations were performed using the Vienna ab initio simulation package (VASP) [23], which solves the Kohn–Sham equation using the plane wave total-energy method to obtain the valence electron density and the wave function [24]. The electron–ion interaction was described by the projector augmented wave (PAW) method. It is worth noting that the narrow f-bands in RE compounds are not adequately described by standard local density approximation (LDA) and the generalized gradient approximation (GGA) [25,26] due to strong electronic correlation effects. In order to deal with these insufficiencies, the DFT + U method [27] is often employed, with U used fit to some spectral data. Another approach is to use a hybrid functional introducing a portion of exact Hartree–Fock type exchange into the exchange–correlation functional [28,29,30]. Both approaches improve the treatment of strongly correlated electrons. However, DFT + U suffers from an ambiguity of the Hubbard U value, while hybrid functionals suffer from extremely demanding computational costs. Generally, the strong correlation of f-electrons have a great influence on the oxide, nitride and insulator systems and they have little influence on the metal system because of the small degree of the electron localization in the metal system [29,31,32,33]. In this study, the Perdew–Burke–Ernzerh (PBE) version of the GGA [34] was used to treat the exchange–correlation functional. In the VASP calculations, geometric optimization was performed by the conjugate gradient algorithm [35] until the total energy changed to <10−6 eV/atom and the Hellmann–Feynman force on the atom [20] was in the range of 10−3 eV/Å. To ensure the calculation accuracy, we performed a series of convergence tests and finally set the truncation energy to 400 eV. The Brillouin zone integral used the high-symmetric k-point method in the form of the Monkhorst–Pack method [36], and the k-point separation was 13 × 13 × 13. The linear tetrahedron method with Blöchl corrections [37] was used for static total energy calculations. The PAW_PBE pseudopotentials for each element in this calculation were chosen as Al, Er_3, Hf_pv, Lu_3, Sc_sv, Si, Ti_sv, Tm_3, Yb_2, Li_sv, Mg, and Zr_sv. Thermal properties were calculated by employing the quasi-harmonic approximation (QHA) theory implemented in the PHONOPY code [38]. For phonon calculations, 3 × 3 × 3 supercells and Monkhorst-Pack 3 × 3 × 3 k-point grids were chosen. At the same time, Hessian matrices (second-order derivative matrices) were determined by the density-generalized perturbation theory, using symmetry to reduce the number of displacements.
The enthalpy, E(T), free energy, F(T), entropy, S(T), and heat capacity, Cυ, can be given as follows [39,40]:
E ( T ) = E t o t + E z p + k B T ω exp ( ω / k B T ) F ( ω ) d ω
F ( T ) = E t o t + E z p + k B T F ( ω ) ln [ 1 exp ( ω / k B T ) ] d ω
S ( T ) = k B { ω / k B T exp ( ω / k B T ) 1 F ( ω ) d ω F ( ω ) ln [ 1 exp ( ω / k B T ) ] d ω }
where T is the temperature, E t o t is the total energy of the system, E z p is the zero-point vibrational energy, is the Planck constant, k B is Boltzmann constant, and F ( ω ) is the phonon density of states.
When the relationship between heat energy and volume is expressed within a quasi-harmonic approximation (QHA), the Helmholtz free energy can be written as follows [41,42]:
F ( V , T ) = U 0 ( V ) + F v i b ( V , T )
where U 0 ( V ) is the zero-temperature classical energy, while
F v i b = 1 2 q , j ω j ( q , V ) + k B T q , j ln [ 1 exp ( ω j ( q , V ) / k B T ) ]
denotes the contribution of vibration to free energy F, ωj(q,V) is the phonon frequency in the j th mode, and q is the wave vector for a given volume V . The coefficient of linear thermal expansion can be expressed as follows:
α ( T ) = 1 3 B q , j γ j ( q ) C v j ( q , T )
where B is the isothermal bulk modulus and   γ j ( q ) is the Grüneisen parameter in the j th mode, which can be expressed as follows:
γ j ( q ) = d [ ln   ω j ( q , V ) ] d ( ln   V )
and C v j ( q , T ) , the mode contribution to the specific heat, is defined as follows:
C v j ( q , T ) = ω j ( q , V )   V d d T [ exp ( ω j ( q , V )   k B T ) 1 ] 1
From the definition of the specific heat, the total specific heat of the crystal can be obtained from the following equation:
C v ( T ) = q , j C v j ( q , T )
The thermal Grüneisen parameter is the weighted average of the mode Grüneisen parameters:
γ th = 1   C v ( T ) q , j γ j ( q ) C v j ( q , T ) = α K T V C v = α K S V C p
where C v and C p denote the heat capacity at a constant volume and pressure, respectively. K T and K S are the isothermal volume modulus and adiabatic volume modulus, respectively. The relationship between C p and C v can be expressed as follows:
C p = C v ( 1 + α γ th T )

3. Results and Discussion

3.1. Thermal Properties

The enthalpy of formation (ΔH) of the ordered L12-Al3M (M = Er, Hf, Lu, Sc, Si, Ti, Tm, Yb, Li, Mg, Zr) phase is defined as the energy gain or loss relative to the unstretched volume. It reflects the alloying capability and thermal stability of the corresponding system [43,44,45] and can be calculated using the following equation [46]:
Δ H ( Al 3 M ) = E ( Al 3 M ) 3 4 E ( Al ) 1 4 E ( M )
where E(Al3M) is the total energy per unit cell of Al3M. E(Al) is the average energy of each Al atom in the FCC structure, and E(M) is the average energy of each atom of Er, Hf, Lu, Mg, Sc, Ti, Tm, and Zr in the hexagonal close-packed structure, Li in the body-centered cubic structure, Si in the diamond structure, and Yb in the FCC structure.
The calculated lattice constants (a0), enthalpy of formation (ΔH), volume (V0) per atom, and lattice mismatch δ of L12 structures at 0 K are listed in Table 1. The lattice mismatch δ can be defined as δ = | a A l 3 M a A l | a A l × 100 % , where a A l 3 M and a A l are the lattice constants of the L12-Al3M and pure Al, respectively. The results show that the lattice mismatches between different L12-structured precipitate phases and Al matrixes are no more than 6.27%. Generally, the lattice mismatch implies a larger stability. Meanwhile, a negative enthalpy of formation reflects that the precipitation process is exothermic, and the smaller the enthalpy of formation value, the more stable the phase. Therefore, the calculated results of lattice mismatches and enthalpy both indicate that the ordered L12 -Al3M phase shows excellent stability in the Al matrix.
Figure 1 shows the temperature dependence of the Gibbs free energy, F(T), and entropy, S(T), of Al and L12-Al3M structures. The graphs in narrower temperature ranges, i.e., 0–400 K and 0–200 K, are presented for better insight. The Gibbs free energy gradually decreases, while the entropy exhibits an increase with the temperature. As illustrated in Figure 1a, the Gibbs free energies of all the L12 structures decrease slowly in the temperature range from 0 to 400 K. In contrast, the decrease is significant in the temperature range of 400–1000 K. Further investigation shows that the doping atoms Yb, Lu, Er, and Tm combine with Al atoms to form L12 structures, decreasing the Gibbs free energy of Al alloys. Remarkably, the Gibbs free energy of Al3Yb is the smallest in the entire temperature range (0–1000 K), indicating that the alloy shows the best thermal stability. However, Hf, Mg, Zr, Li, Sc and Ti atoms do not improve the stability of the Al alloys. The Gibbs free energies of the Al3M phases varied with temperature in the following order: Al3Yb < Al3Lu < Al3Er < Al3Tm < Al < Al3Hf < Al3Mg < Al3Zr < Al3Li < Al3Sc < Al3Ti, which shows that the Yb, Lu, Er and Tm elements would improve the thermal stability of Al alloys.
The entropy of the system exhibits a distinct behavior. Entropy is considered to be a measure of disorder, and it never decreases because isolated systems spontaneously evolve toward thermodynamic equilibrium, which is the state of maximum entropy (minimum energy). As shown in Figure 1b, the entropy significantly increases with the temperature in the range of 0–400 K, while its growth rate gradually slows down at higher temperatures. Some of the L12 phases, i.e., Al3M (M = Yb, Lu, Er, and Tm), increase the entropy of the Al alloy system at different temperatures. The slope of the curve for Al3Yb is the largest in the entire temperature range, implying that the Al3Yb structure exhibits the highest thermal stability. The change of entropy of the Al3M phases with temperature is arranged in the following order: Al3Yb > Al3Lu > Al3Er > Al3Tm > Al > Al3Hf > Al3Ti > Al3Mg > Al3Zr > Al3Li > Al3Sc.
The temperature-dependent isobaric heat capacity (Cp) curves of Al and L12-Al3M structures and the corresponding dependence in a narrower temperature range are shown in Figure 2. The Cp values of Al and Al3M increase dramatically with temperature in the range of 0–300 K, and the rate increases gradually and the Cp value finally converges to 100 J/K/mol at about 700 K. Compared with Al, the Cp value of some Al3M (M = Yb, Lu, Er, and Tm) structures increases significantly, indicating a stronger thermal stability than for bulk Al. The Al3Yb alloy firstly reaches a constant value of 100 J/K/mol at 0–500 K, confirming the beneficial effect of Yb atoms on the thermal stability of the Al alloys. However, Ti and Mg atoms do not show any positive effect on the thermal stability of Al alloys. Lu, Er and Tm atoms have a positive and significant effect on the thermal stability of the Al alloys, especially in the temperature range of 0–600 K. The temperature-dependent relationship of the Cp of Al and its alloys is arranged in the following order: Al3Yb > Al3Lu > Al3Er > Al3Tm > Al4 > Al3Mg > Al3Hf > Al3Zr > Al3Sc > Al3Li > Al3Ti.
Remarkably, the descending order of Gibbs free energy is consistent with the ascending order of entropy and Cp, especially for Al3Yb, Al3Lu, Al3Er, and Al3Tm alloys. Considering that those structures have no enhancing effect on the thermodynamic properties of Al alloys, we focus on Al3Yb, Al3Lu, Al3Er, and Al3Tm structures.
In Figure 3a, we present the temperature-dependent coefficient of thermal expansion (α) of L12 structures calculated using Equation (6). The α value sharply increases with the temperature from 0 to 300 K and then gradually slows down. Above 600 K, the thermal expansion coefficient of the compound reaches a plateau, meaning that the material produces a linear thermal expansion. As described in thermal theory, the temperature dependence of α(T) is very similar to that of the isobaric heat capacity Cp(T). In the temperature range from 0 to 125K, the α value of Al3Yb is the largest, but at 125 K, α-Al3Mg reaches the highest α value, indicating that Al3Mg is more prone to expansion at a certain temperature. This behavior is consistent with the trend of their Grüneisen parameters with the temperature increase. A high temperature increases the thermal effect on the lattice constant, raising the thermal expansion coefficient.
The Grüneisen parameter (γ) is a key physical quantity for the thermoelasticity of materials [52,53,54]. At the same time, it is the basis for explaining phonon frequency variation with volume [55,56]. Figure 3b shows the temperature dependence curves of the Grüneisen parameter defined by Equation (10) for the L12 structures. The Al3M phases exhibit the different temperature dependence of the Grüneisen parameter, γ, in the low-temperature region from 0 to 200 K. For Al3Sc, Al3Hf, and Al3Zr, the γ value increases with the temperature, while for other compounds it decreases firstly in the low temperature range, and then increases. As we know, due to the quantum effect, the lattice and electron vibration of materials is more complex at low temperatures. For example, in the low-temperature limit, as the frequency of the excited lattice modes goes to zero, the volume dependences of the frequencies, are related to the volume dependences of the corresponding elastic constants [57,58]. This means that the variation law of the Grüneisen parameter at a low temperature is rather complicated, and the elastic-related properties remain to be calculated, which will not be discussed in this study. As expected, the γ values of all the alloys converge to a certain constant above 300 K. It can be found that in the temperature range of 100–1000 K, the Grüneisen parameter is sorted out as Al3Yb > Al3Lu > Al3Er > Al3Tm, which is in good agreement with the thermodynamic stability order discussed earlier.
In summary, Yb, Lu, Er, and Tm atoms with bigger atomic radii than the Al atom are beneficial for improving the thermal properties of Al Alloys, and the thermal stability of the L12-Al3M phases is arranged in the order of Al3Yb > Al3Lu > Al3Er > Al3Tm> Al. We would like to take the relationship between entropy and atomic radius at 1000K as an example to show the obvious relationship between the thermodynamic properties of the L12-Al3M structures and the atomic radius of alloying atoms M, and the result is illustrated in Figure 4. It is obvious that the structures are of L12-Al3M (M = Yb, Lu, Er, and Tm) with better thermal properties and a bigger radius in the M atom. The precise atomic radius values of Yb, Lu, Er, and Tm atoms are 1.940, 1.734, 1.757, and 1.746, respectively, which are are more than 0.2 times larger than Al.

3.2. Electronic Structure

The microstructure of materials is closely correlated to their electron structures [59,60]. When alloying atoms join in an Al matrix, they inevitably cause the electron charge redistribution. Hence, the underlying mechanisms of the change in thermal properties can be analyzed and understood in terms of charge density redistribution. Differential charge density can clearly give information about inter-atomic interactions, such as charge transfer and bonding types between atoms. Bonding charge density (BCD) is defined as a difference in the self-consistent charge density, ρ, of the compound from the first-principle calculations and the reference charge density constructed from the simple superposition of the non-interacting atomic charge density at the crystal sites. The differential charge density, Δρ, is defined as Δρ = ρ(Al3M) − ρ(Al atoms) − ρ(M atoms).
Figure 5 displays a 3D plot of the differential charge density of the L12-Al3M compounds. The color of the electron cloud indicates the degree of charge enrichment, with red and green indicating the charge gain and reduction, respectively. Most of the electron clouds are localized between two adjacent atoms of Al and M, showing covalent bonding properties, but the bonding strength varies between the different atom types. At the same time, the electron cloud of Al3Er, Al3Lu, and Al3Tm alloys are mainly localized at the tetrahedral sites. In contrast, the charge density is significantly reduced at the octahedral sites, mainly due to the hybridization of d or f orbitals of M atoms with their nearest neighbors. Conclusively, these L12-Al3M structures have good thermal stability due to the strong covalent bonding between Al-M atoms.
Figure 6 and Figure 7 describe the charge density distribution of bonding on the {100} and {111} surfaces of the L12-Al3M phases, respectively. The color shades reflect the strength of the bonding energy: red indicates an increase in the charge accumulation after bonding (i.e., the increase in electron density); blue represents a decrease in the charge density. Electrons are localized between the Al and M atoms, showing strong covalent bonding. From Figure 6 it can be seen that the doping with M elements forms strong Al-M bonds (M = Er, Tm, Yb, and Lu). Figure 7 shows that the charge decreases around the Al and M atoms, while there is a strong tendency of charge accumulation between them, which corroborates well the results shown in Figure 5.
At the same time, we performed a Bader charge analysis on these structures to quantify the amount of charge transfer, which is the most intuitive and simplest way to obtain the charge transfer number between atoms. Table 2 summarizes the Bader charge analysis of several doped elements of the L12-Al3M structure. A large amount of charge migrates from the doping element M to Al atoms, resulting in strong Al-M bonds, which provide a strong intrinsic basis for the stability of the L12-Al3M structures.

4. Conclusions

We systematically analysed the therodynamic properties of L12-Al3M (M = Er, Hf, Lu, Sc, Si, Ti, Tm, Yb, Li, Mg, Zr) compounds using density functional theory. The small enthalpies and lattice mismatches indicate that the L12-Al3X (M = Er, Hf, Lu, Sc, Ti, Tm, Yb, Li, Mg, Zr) structural compounds could exist stably in an Al matrix. Furthermore, the thermodynamic data (including the Gibbs free energy F, entropy S, isobaric heat capacity Cp, thermal expansion coefficient α and Grüneisen parameter γ) implied that the addition of Yb, Lu, Er, and Tm atoms with an atomic radius 0.2 times lager than Al has a positive effect on the thermal stability of the Al alloys. The thermal stability of these structures is ranked in the following order: Al3Yb > Al3Lu > Al3Er > Al3Tm > Al. The internal cause of this thermal stability was investigated from the microscopic point of view of electron bonding through the analysis of charge density. A differential charge density map and Bader charge analysis show that covalent bonds are easily formed between M (M = Er, Lu, Tm, Yb) and Al. The charge density is mainly distributed in the <111> direction at the tetrahedral sites, while the charge density is significantly reduced at the octahedral sites. The strong bonding of the Al-M bonds provides a strong intrinsic basis for the stability of the Al3M structures.
It is found that the elements Er, Lu, Tm, and Yb are suitable replacements of Sc, which would reduce the cost and improve the thermodynamic properties of aluminum alloys at the same time.

Author Contributions

Conceptualization and methodology, L.L., M.H., J.L. (Jianbo Li), X.P.; software, T.F.; investigation, Z.C., Q.Z.; validation and formal analysis, J.L. (Jihang Lan); data curation, writing—original draft preparation, (Jihang Lan); review and editing, L.L.; visualization, J.L. (Jihang Lan); supervision, L.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by the National Natural Science Foundation of China (Nos. 51701243), the Hunan Provincial Natural Science Foundation of China (No. 2019JJ40544), the excellent youth funding of Hunan Provincial Education Department (No.19B587) and the PhD Research Startup Foundation of Central South University of Forestry and Technology (No. 2017YJ020), as well as by the R & D plan for key areas in Guangdong Province (2020B010186001) and Special Innovation Projects of Universities in Guangdong Province (2018KTSCX240).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data that support the findings of this study are available on request from the corresponding author. The data are not publicly available due to privacy or ethical restrictions.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ye, H. An overview of the development of Al-Si-alloy based material for engine applications. J. Mater. Eng. Perform. 2003, 12, 288–297. [Google Scholar] [CrossRef]
  2. Cheng, S.; Zhao, Y.H.; Zhu, Y.T.; Ma, E. Optimizing the strength and ductility of fine structured 2024 Al alloy by nano-precipitation. Acta Mater. 2007, 55, 5822–5832. [Google Scholar] [CrossRef]
  3. Das, S. Development of aluminium alloy composites for engineering applications. Trans. Indian Inst. Met. 2004, 57, 325–334. [Google Scholar]
  4. Knipling, K.E.; Dunand, D.C.; Seidman, D.N. Criteria for developing castable, creep-resistant aluminum-based alloys—A review. J. Alloys Compd. 2006, 97, 246–265. [Google Scholar] [CrossRef] [Green Version]
  5. Mishra, R.S.; Ma, Z.Y. Friction stir welding and processing. Mater. Sci. Eng. R 2005, 50, 1–78. [Google Scholar] [CrossRef]
  6. Wang, G.; Zhao, Y.; Hao, Y. Friction stir welding of high-strength aerospace aluminum alloy and application in rocket tank manufacturing. J. Mater. Sci. Technol. 2018, 34, 73–91. [Google Scholar] [CrossRef]
  7. Duan, Y.H.; Sun, Y.; Peng, M.J.; Zhou, S.G. Ab-initio investigations on elastic properties in L12 structure Al3Sc and Al3Y under high pressure. J. Alloys Compd. 2014, 585, 587–593. [Google Scholar] [CrossRef]
  8. Zhang, X.; Jiang, W. First-principles investigations on vibrational, thermodynamic, mechanical properties and thermal conductivity of L12 Al3X (X = Sc, Er, Tm, Yb) intermetallics. Phys. Scripta 2015, 90, 065701. [Google Scholar] [CrossRef]
  9. Zhang, Y.; Gao, K.; Wen, S.; Huang, H.; Nie, Z.; Zhou, D. The study on the coarsening process and precipitation strengthening of Al3Er precipitate in Al–Er binary alloy. J. Alloys Compd. 2014, 610, 27–34. [Google Scholar] [CrossRef]
  10. Jahnátek, M.; Krajčí, M.; Hafner, J. Response of trialuminides to [110] uniaxial loading: An ab initio study for Al3(Sc, Ti, V). Phys. Rev. B 2007, 76, 014110. [Google Scholar] [CrossRef]
  11. 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]
  12. Watanabe, C.; Kondo, T.; Monzen, R. Coarsening of Al3Sc precipitates in an Al-0.28 wt pct Sc alloy. Metall. Mater. Trans. A 2004, 35, 3003–3008. [Google Scholar] [CrossRef]
  13. Karnesky, R.A.; Van Dalen, M.E.; Dunand, D.C.; Seidman, D.N. Effects of substituting rare-earth elements for scandium in a precipitation-strengthened Al–0.08 at.% Sc alloy. Scripta Mater. 2006, 55, 437–440. [Google Scholar] [CrossRef]
  14. Van Dalen, M.E.; Karnesky, R.A.; Cabotaje, J.R.; Dunand, D.C.; Seidman, D.N. Erbium and ytterbium solubilities and diffusivities in aluminum as determined by nanoscale characterization of precipitates. Acta Mater. 2009, 57, 4081–4089. [Google Scholar] [CrossRef]
  15. Mao, Z.; Seidman, D.N.; Wolverton, C.J.A.M. First-principles phase stability, magnetic properties and solubility in aluminum–rare-earth (Al–RE) alloys and compounds. Acta Mater. 2011, 59, 3659–3666. [Google Scholar] [CrossRef]
  16. Peng, G.; Chen, K.; Fang, H.; Chen, S. Effect of Cr and Yb additions on microstructure and properties of low copper Al–Zn–Mg–Cu–Zr alloy. Mater. Des. (1980–2015) 2012, 36, 279–283. [Google Scholar] [CrossRef]
  17. Czerwinski, F. Thermal stability of aluminum alloys. Materials 2020, 13, 3441. [Google Scholar] [CrossRef] [PubMed]
  18. Zhang, X.D.; Jiang, W. Lattice stabilities, mechanical and thermodynamic properties of Al3Tm and Al3Lu intermetallics under high pressure from first-principles calculations. Chin. Phys. B. 2015, 25, 026301. [Google Scholar] [CrossRef]
  19. Li, D.L.; Chen, P.; Yi, J.X.; Tang, B.Y.; Peng, L.M.; Ding, W.J. Ab initio study on the thermal properties of the fcc Al3Mg and Al3Sc alloys. J. Phys. D Appl. Phys. 2009, 42, 225407. [Google Scholar] [CrossRef]
  20. Pan, R.K.; Wang, H.C.; Shi, T.T.; Tian, X.; Tang, B.Y. Thermal properties and thermoelasticity of L12 ordered Al3RE (RE= Er, Tm, Yb, Lu) phases: A first-principles study. Mater. Des. 2016, 102, 100–105. [Google Scholar] [CrossRef]
  21. Kerkove, M.A.; Wood, T.D.; Sanders, P.G.; Kampe, S.L.; Swenson, D. The diffusion coefficient of scandium in dilute aluminum-scandium alloys. Metall. Mater. Trans. A 2014, 45, 3800–3805. [Google Scholar] [CrossRef]
  22. Rokhlin, L.L.; Bochvar, N.R.; Dobatkina, T.V.; Leont’ev, V.G. Al-rich portion of the Al-Hf phase diagram. Russ. Metall. (Met.) 2009, 3, 258–262. [Google Scholar] [CrossRef]
  23. 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]
  24. Arya, A.; Carter, E.A. Structure, bonding, and adhesion at the TiC (100)/Fe (110) interface from first principles. J. Chem. Phys. 2003, 118, 8982–8996. [Google Scholar] [CrossRef] [Green Version]
  25. Kohn, W.; Sham, L.J. Self-consistent equations including exchange and correlation effects. Phys. Rev. 1965, 140, A1133. [Google Scholar] [CrossRef] [Green Version]
  26. Perdew, J.P.; Wang, Y. Erratum: Accurate and simple analytic representation of the electron-gas correlation energy. [Phys. Rev. B 45, 13244 (1992)]. Phys. Rev. B 2018, 98, 079904. [Google Scholar] [CrossRef] [Green Version]
  27. Anisimov, V.I.; Zaanen, J.; Andersen, O.K. Band theory and Mott insulators: Hubbard U instead of Stoner, I. Phys. Rev. B 1991, 44, 943. [Google Scholar] [CrossRef] [Green Version]
  28. Heyd, J.; Scuseria, G.E.; Ernzerhof, M. Hybrid functionals based on a screened Coulomb potential. J. Chem. Phys. 2003, 118, 8207–8215. [Google Scholar] [CrossRef] [Green Version]
  29. Da Silva, J.L.; Ganduglia-Pirovano, M.V.; Sauer, J.; Bayer, V.; Kresse, G. Hybrid functionals applied to rare-earth oxides: The example of ceria. Phys. Rev. B 2007, 75, 045121. [Google Scholar] [CrossRef]
  30. Hay, P.J.; Martin, R.L.; Uddin, J.; Scuseria, G.E. Theoretical study of CeO2 and Ce2O3 using a screened hybrid density functional. J. Chem. Phys. 2006, 125, 034712. [Google Scholar] [CrossRef]
  31. Topsakal, M.; Wentzcovitch, R.M. Accurate projected augmented wave (PAW) datasets for rare-earth elements (RE= La–Lu). Comp. Mater. Sci. 2014, 95, 263–270. [Google Scholar] [CrossRef] [Green Version]
  32. Beiranvand, R. Hybrid exchange–correlation energy functionals for accurate prediction of the electronic and optical properties of alkaline-earth metal oxides. Mater. Sci. Semicond. Process. 2021, 135, 106092. [Google Scholar] [CrossRef]
  33. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar] [CrossRef] [Green Version]
  34. Gonze, X. Towards a potential-based conjugate gradient algorithm for order-N self-consistent total energy calculations. Phys. Rev. B 1996, 54, 4383. [Google Scholar] [CrossRef] [PubMed]
  35. Feynman, R.P. Forces in molecules. Phys. Rev. 1939, 56, 340. [Google Scholar] [CrossRef]
  36. Monkhorst, H.J.; Pack, J.D. Special points for Brillouin-zone integrations. Phys. Rev. B. 1976, 13, 5188. [Google Scholar] [CrossRef]
  37. Blöchl, P.E.; Jepsen, O.; Andersen, O.K. Improved tetrahedron method for Brillouin-zone integrations. Phys. Rev. B. 1994, 49, 16223. [Google Scholar] [CrossRef]
  38. Togo, A.; Tanaka, I. First principles phonon calculations in materials science. Scr. Mater. 2015, 108, 1–5. [Google Scholar] [CrossRef] [Green Version]
  39. Baroni, S.; De Gironcoli, S.; Dal Corso, A.; Giannozzi, P. Phonons and related crystal properties from density-functional perturbation theory. Rev. Mod. Phys. 2001, 73, 515. [Google Scholar] [CrossRef] [Green Version]
  40. Jha, P.K. Phonon spectra and vibrational mode instability of MgCNi3. Phys. Rev. B 2005, 72, 214502. [Google Scholar] [CrossRef]
  41. Niessen, A.D.; De Boer, F.R.; Boom, R.D.; De Chatel, P.F.; Mattens, W.C.M.; Miedema, A.R. Model predictions for the enthalpy of formation of transition metal alloys II. Calphad 1983, 7, 51–70. [Google Scholar] [CrossRef]
  42. Hu, W.C.; Liu, Y.; Li, D.J.; Zeng, X.Q.; Xu, C.S. Mechanical and thermodynamic properties of Al3Sc and Al3Li precipitates in Al–Li–Sc alloys from first-principles calculations. Physica B 2013, 427, 85–90. [Google Scholar] [CrossRef]
  43. 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]
  44. 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]
  45. Gayle, F.W.; Vandersande, B. Phase transformations in the Al-Li-Zr system. Acta Metall. 1989, 37, 1033–1046. [Google Scholar] [CrossRef]
  46. Hu, Z.; Yin, Z.; Lin, J.; Zhang, L.; Qiu, S.; Yan, H.; Song, H. Microstructural, electronic, and mechanical properties of L12 ordered Al3Er and Al3Yb intermetallics: An experimental and first-principles calculations. Mater. Res. Express 2019, 6, 116516. [Google Scholar] [CrossRef]
  47. Dean, J.A. Lange’s Handbook of Chemistry. Mater. Manuf. Process 1990, 5, 687–688. [Google Scholar] [CrossRef]
  48. Li, S.S.; Li, L.; Han, J.; Wang, C.T.; Xiao, Y.Q.; Jian, X.D.; 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]
  49. Tao, X.; Ouyang, Y.; Liu, H.; Feng, Y.; Du, Y.; Jin, Z. First-principles calculations of the thermodynamic and elastic properties of the L12-based Al3RE (RE= Sc, Y., La–Lu). Int. J. Mater. Res. 2008, 99, 582–588. [Google Scholar] [CrossRef]
  50. Ghosh, G.; Asta, M. First-principles calculation of structural energetics of Al–TM (TM= Ti, Zr, Hf) intermetallics. Acta Mater. 2005, 53, 3225–3252. [Google Scholar] [CrossRef]
  51. Wolverton, C. Crystal structure and stability of complex precipitate phases in Al–Cu–Mg–(Si) and Al–Zn–Mg alloys. Acta Mater. 2001, 49, 3129–3142. [Google Scholar] [CrossRef]
  52. Wood, I.G.; Vočadlo, L.; Dobson, D.P.; Price, G.D.; Fortes, A.D.; Cooper, F.J.; McCammon, C.A. Thermoelastic properties of magnesiowüstite,(Mg1−xFex)O: Determination of the Anderson–Grüneisen parameter by time-of-flight neutron powder diffraction at simultaneous high pressures and temperatures. J. Appl. Crystallogr. 2008, 41, 886–896. [Google Scholar] [CrossRef]
  53. Yang, H.; Ghose, S. Thermal expansion, Debye temperature and Grüneisen parameter of synthetic (Fe, Mg) SiO3 orthopyroxenes. Phys. Chem. Miner. 1994, 20, 575–586. [Google Scholar] [CrossRef]
  54. Petrova, E.; Ermilov, S.; Su, R.; Nadvoretskiy, V.; Conjusteau, A.; Oraevsky, A. Using optoacoustic imaging for measuring the temperature dependence of Grüneisen parameter in optically absorbing solutions. Opt. Express 2013, 21, 25077–25090. [Google Scholar] [CrossRef] [Green Version]
  55. Kushwah, S.S.; Tomar, Y.S.; Upadhyay, A.K. On the volume-dependence of the Grüneisen parameter and the Lindemann law of melting. J. Phys. Chem. Solids 2013, 74, 1143–1145. [Google Scholar] [CrossRef]
  56. Kushwah, S.S.; Sharma, M.P. Volume dependence of the Grüneisen parameter for MgO. Solid State Commun. 2012, 152, 414–416. [Google Scholar] [CrossRef]
  57. Adesakin, G.; Olubosede, O.; Fatigun, A.; Aliyu, E.; Oyedele, E.; Adekoya, O.E.; Ewumi10, T.O. Gruneisen parameter of metals based on free electron theory. Int. J. Mech. Prod. Eng. Res. Dev. (IJMPERD) 2020, 10, 5107–5114. [Google Scholar]
  58. Smith, T.F.; White, G.K. The low-temperature thermal expansion and Grüneisen parameters of some tetrahedrally bonded solids. J. Phys. C Solid State Phys. 1975, 8, 2031. [Google Scholar] [CrossRef]
  59. Wang, Z.; Chen, D.; Fang, Q.; Chen, H.; Fan, T.; Liu, B.; Tang, P. Effects of finite temperature on the surface energy in Al alloys from first-principles calculations. Appl. Surf. Sci. 2019, 479, 499–505. [Google Scholar] [CrossRef]
  60. Tian, T.; Wang, X.F.; Li, W. Ab initio calculations on elastic properties in L12 structure Al3X and X3Al-type (X= transition or main group metal) intermetallic compounds. Solid State Commun. 2013, 156, 69–75. [Google Scholar] [CrossRef]
Figure 1. The calculated temperature dependence of (a) the Gibbs free energy and (c) the entropy of pure Al and L12-Al3M structures, where (b,d) represent this dependence, are limited to the specific temperature range of 0–400 K and 0–200 K, respectively.
Figure 1. The calculated temperature dependence of (a) the Gibbs free energy and (c) the entropy of pure Al and L12-Al3M structures, where (b,d) represent this dependence, are limited to the specific temperature range of 0–400 K and 0–200 K, respectively.
Metals 11 01420 g001
Figure 2. (a) The isobaric heat capacity Cp–temperature curves of pure Al and L12-Al3M structures; (b) Cp in a narrower temperature range of 0–400 K.
Figure 2. (a) The isobaric heat capacity Cp–temperature curves of pure Al and L12-Al3M structures; (b) Cp in a narrower temperature range of 0–400 K.
Metals 11 01420 g002
Figure 3. (a) Coefficient of thermal expansion α, and (b) Grüneisen parameter,γ for Al and L12-Al3M phases as a function of temperature.
Figure 3. (a) Coefficient of thermal expansion α, and (b) Grüneisen parameter,γ for Al and L12-Al3M phases as a function of temperature.
Metals 11 01420 g003
Figure 4. The relationship between entropy and solute atomic (M) radius of Al3M at 1000 K.
Figure 4. The relationship between entropy and solute atomic (M) radius of Al3M at 1000 K.
Metals 11 01420 g004
Figure 5. 3D map of the differential charge density of L12-Al3M (M = Yb, Lu, Er, and Tm) compounds.
Figure 5. 3D map of the differential charge density of L12-Al3M (M = Yb, Lu, Er, and Tm) compounds.
Metals 11 01420 g005
Figure 6. Charge density distribution of the L12-Al3M phases in (100) plane: (a) Al3Yb, (b) Al3Lu, (c) Al3Er, and (d) Al3Tm. All the contour plots are on the same scale, and the interval between the contour lines is 0.001e/bohr3.
Figure 6. Charge density distribution of the L12-Al3M phases in (100) plane: (a) Al3Yb, (b) Al3Lu, (c) Al3Er, and (d) Al3Tm. All the contour plots are on the same scale, and the interval between the contour lines is 0.001e/bohr3.
Metals 11 01420 g006
Figure 7. Charge density distribution of the L12-Al3M phase in (111) plane: (a) Al3Yb, (b) Al3Lu, (c) Al3Er, and (d) Al3Tm. All the contour plots have the same scale, and the interval between the contour lines is 0.001e/bohr3.
Figure 7. Charge density distribution of the L12-Al3M phase in (111) plane: (a) Al3Yb, (b) Al3Lu, (c) Al3Er, and (d) Al3Tm. All the contour plots have the same scale, and the interval between the contour lines is 0.001e/bohr3.
Metals 11 01420 g007
Table 1. The Solute atomic radius (Å) and calculated lattice constants, a0 (Å), formation enthalpy, ΔH (kJ·mol−1), volume per atom, V03 atom−1), and lattice mismatch, δ (%), of L12 structures at 0 K.
Table 1. The Solute atomic radius (Å) and calculated lattice constants, a0 (Å), formation enthalpy, ΔH (kJ·mol−1), volume per atom, V03 atom−1), and lattice mismatch, δ (%), of L12 structures at 0 K.
StructuresSolute Atomic Radius (Å) Lattice   Constants ,   a 0   ( Å ) ΔH (kJ·mol−1)V0 (Å3 atom−1)δ (%)
This workLiteratureThis workLiterature
Al bulk1.431 [47]4.0384.039 [48]16.474
Al3Er1.757 [47]4.2324.215 [49]−39.949−40.000 [49]18.9504.80
Al3Hf1.564 [47]4.0904.081 [50]−35.341−37.300 [50]17.1081.29
Al3Li1.520 [47]4.0254.025 [48]−9.452−9.439 [48]16.3080.32
Al3Lu1.734 [47]4.2694.206 [49]−38.916−38.920 [49]19.4535.72
Al3Mg1.600 [47]4.1384.146 [51]−0.898−1.060 [51]17.7102.48
Al3Sc1.630 [47]4.1034.106 [48]−43.649−43.826 [48]17.3081.61
Al3Ti1.448 [47]3.9783.977 [48]−35.422−34.579 [48]15.7351.49
Al3Tm1.746 [47]4.2214.200 [49]−39.321−39.000 [49]18.8054.53
Al3Yb1.940 [47]4.2914.200 [49]−16.312−17.000 [49]19.7536.27
Al3Zr1.600 [47]4.1074.106 [48]−44.210−44.018 [48]17.3231.71
Table 2. The calculated charge transfer numbers caused by solute atoms in the L12-Al3M structures.
Table 2. The calculated charge transfer numbers caused by solute atoms in the L12-Al3M structures.
Solutes/SolventYbLuErTm
−0.7329−1.0468−0.9388−0.9579
Al10.33140.25930.2770 0.4411
Al20.21310.38920.3889 0.2740
Al30.18850.39830.2729 0.2429
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lan, J.; Chen, Z.; Liu, L.; Zhang, Q.; He, M.; Li, J.; Peng, X.; Fan, T. The Thermal Properties of L12 Phases in Aluminum Enhanced by Alloying Elements. Metals 2021, 11, 1420. https://doi.org/10.3390/met11091420

AMA Style

Lan J, Chen Z, Liu L, Zhang Q, He M, Li J, Peng X, Fan T. The Thermal Properties of L12 Phases in Aluminum Enhanced by Alloying Elements. Metals. 2021; 11(9):1420. https://doi.org/10.3390/met11091420

Chicago/Turabian Style

Lan, Jihang, Zhaoqun Chen, Linghong Liu, Qingzhou Zhang, Mengdong He, Jianbo Li, Xiaofang Peng, and Touwen Fan. 2021. "The Thermal Properties of L12 Phases in Aluminum Enhanced by Alloying Elements" Metals 11, no. 9: 1420. https://doi.org/10.3390/met11091420

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