Next Article in Journal
Tuning the Hardness of Produced Parts by Adjusting the Cooling Rate during Laser-Based Powder Bed Fusion of AlSi10Mg by Adapting the Process Parameters
Next Article in Special Issue
FEM Simulation on First-Step Drawing Process of Platinum-Clad Nickel Bars
Previous Article in Journal
Hydriding, Oxidation, and Ductility Evaluation of Cr-Coated Zircaloy-4 Tubing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Alloying Strategy to Tailor the Mechanical Properties of θ-Al13Fe4 Phase in Al-Mg-Fe Alloy by First-Principles Calculations

National Engineering Research Centre for Advanced Titanium and Titanium Alloy Material Technology, Luoyang Ship Material Research Institute, Luoyang 471023, China
*
Author to whom correspondence should be addressed.
Metals 2022, 12(12), 1999; https://doi.org/10.3390/met12121999
Submission received: 28 October 2022 / Revised: 13 November 2022 / Accepted: 17 November 2022 / Published: 22 November 2022

Abstract

:
As an important strengthening phase in Al-Mg-Fe alloy, the elastic and ductile–brittle characteristics of Al13Fe4 intermetallics hold prime significance in ascertaining the mechanical properties and potential application of Al-Mg-Fe alloys. In this study, multialloying of Co, Cu, Cr, Mn, and Ni has been adopted for tuning the mechanical characteristics of the Al13Fe4 phase; their effects on mechanical features and electronic structure of the Al13Fe4 phase have been scrutinized systematically by first-principles calculations employing the density functional theory. The replacement of Fe with M (M = Co, Cu, Cr, Mn, and Ni) is energetically advantageous at 0 K, as evidenced by the negative cohesive energy and mixing enthalpy of all Al13(Fe,M)4 phases. Cu and Ni, on the contrary, have a detrimental impact on Al13Fe4′s modulus and hardness due to the evolution of chemical bonding strength. Co, Cr, and Mn are thus, interesting candidate elements. In the light of B/G and Poisson’s ratio (σ) criteria, Al13Fe4, Al13(Fe,Cu)4, and Al13(Fe,Ni)4 have superior ductility; however, Al13(Fe,Co), Al13(Fe,Mn), and Al13(Fe,Cr)4 tend to be brittle materials. Calculation-based findings show that Co, Cr, and Mn are appropriate alloying elements for enhancing fracture toughness, whereas Mn reduces Al13Fe4′s elastic anisotropy. The electronic structure assessment found that the mechanical properties of the intermetallics are predominantly influenced by the Al-M bonds when the alloying element M replaced Fe.

1. Introduction

During the solidification of aluminium, iron usually precipitates as detrimental Fe-containing intermetallics which can be attributed to the low solubility of Fe in solid α-Al [1,2,3]. The undesired Fe-containing intermetallics that might surface in the course of solidification of Al-Mg alloys are θ-Al13Fe4, Alm(FeMn), Al3(FeMn), Al6(FeMn), β-Al5FeSi (β-Fe) and α-Al15 (FeMn)3Si2 (α-Fe), [4]. Needle-like or plate-like Fe-containing intermetallics, which are brittle, often serve as stress concentration points, leading to crack initiation, propagation, and fracture of alloys [5,6,7,8,9]. To mitigate the deleterious effects of Fe in Al oalloys, several approaches have been utilized, including element neutralization, Fe-containing intermetallics separation, and improved melt cooling rate [10,11,12]. Studies have shown that increasing the solidification rate and the inclusion of Mn, V and Co can prevent plate-like intermetallics from forming and boost alloy yield strength [13,14]. Additionally, it was discovered that Fe, when it exists in the equilibrium form of the θ-Al13Fe4 phase, can be used a3s a h4elpful entity to enhance the mechanical properties of alloys comprising aluminium, magnesium, and manganese [15]. This discovery paves the way for developing a new strategy for enhancing the mechanical properties of secondary alloys made of aluminium, magnesium, and manganese via the generation of an Al-AlFe eutectic microstructure [15]. When Fe content is less than 2.0 wt%, Zhu et al. suggested that the creation of eutectic θ-Al13Fe4 can increase the ultimate tensile strength of Al-Mg-Mn-Fe. The ultimate tensile strength (UTS) and yield strength (YS) of recycled Al-Mg-Mn-Fe alloys underwent an increase from 146 MPa to 289 MPa and 122 MPa to 244 Mpa, respectively, according to Zhao et al.’s research. This is because of the blockage of slip lines by the Fe-rich phases [15]. Al-Mg-Fe alloys’ ability to further improve their mechanical characteristics is constrained by the inherent brittleness of the -Al13Fe4 phase and this phase’s mechanical and elastic anisotropy has a significant impact on the applications for these alloys [16].
In Fe-containing intermetallics, alloying elements such as Mn, Cr, Si, Ni, and Cu may be introduced [17], which do not only modify the crystal structure but also alter the lattice parameters, thereby changing the overall mechanical characteristics [18]. This offers a new set of guidelines for the tuning and optimization of the Al13Fe4 phase’s mechanical characteristics and elastic anisotropy. Numerous researchers have undertaken investigations on the composition design and characteristics optimization of intermetallics that contain iron (Fe) [19,20,21]. According to Chen et al. estimations of the elastic properties of FeAl intermetallics with V, Cr, or Ni doping, these alloying elements were reported to increase the ductility of FeAl alloys, which is presumably because of the reduction in directional chemical bonding in FeAl intermetallics, following V, Cr, and Ni doping. Density functional theory (DFT) simulations were used by Fang et al. to explore Si solution in θ-Al13Fe4 in a systematic manner. The results showed that Si doping in θ-Al13Fe4 will result in greater Si-Fe chemical bonding, thereby leading to a change in the phase’s mechanical properties.
However, it is still unclear as to how exactly the alloying metals such as Co, Cu, Cr, Mn, and Ni affect the mechanical characteristics and elastic anisotropy of θ-Al13Fe4. At present, there is a lack of experimental results on the elastic properties of θ-Al13Fe4, and the study of θ-Al13Fe4 phase does not consider the effect of substitution of these alloying elements on its mechanical properties. For the design and optimization of the mechanical properties of the Al13Fe4 phase, the underlying mechanism must be clearly understood. A trustworthy and effective method for forecasting the stability, electrical structure, and mechanical characteristics of diverse intermetallic and alloy systems is a first-principles computation based on density functional theory (DFT) [22].
In the current work, first-principles calculations employing DFT were employed to examine the impact of Co, Cu, Cr, Mn, and Ni doping upon the stability, mechanical properties, and electronic structures of θ-Al13Fe4 phases. This work investigates the mixing enthalpy, Mulliken’s bond population, the spin-polarized density of states (spin-DOS), and mechanical characteristics for understanding the alteration in the mechanical and physical properties of the Co, Cu, Cr, Mn, and Ni doping in θ-Al13Fe4.

2. Experiments Methods and Details of Calculations

2.1. Experimental Methods

Pure Al (99.99%), pure Mg (99.99%), and master alloys of Al-20Fe are used to create the experimental alloys (all mentioned as mass fractions hereafter, unless specified otherwise). By using a direct-reading spectrometer (Spectrometer LAB, Ametek, kleve, Germany), the experimental chemical compositions of the alloys (as they were made) were determined and are displayed in Table 1. In an electromagnetic induction furnace, the experimental alloys were first melted for 30 min at 750 °C. After that, they were transported to a steel mould, where they are allowed to cool gradually at a rate of 5 °C/min. The alloys are then subjected to composition and microstructure analysis.
Samples are mechanically ground first for microstructural analysis and then polished with OP-U colloidal silica suspension having a grain size of 50 nm. With the aid of a scanning electron microscope (SEM, phenom XI, Phenom-Scientific, The Netherlands), the microstructures of the alloys are examined, and an energy-dispersive spectrometer (EDS) coupled with an SEM is used to ascertain the composition of primary θ-Al13Fe4. The transmission electron microscopy (TEM) method is used to characterize the high-resolution lattice and atomic pictures of the θ-Al13Fe4 phase (TEM, JEM-2100, JEOL, Tokyo, Japan). The ion thinning device (Gatan 691) is used to treat the samples for TEM to ensure that the thickness is less than 10 nm.

2.2. Calculation Details

First-principles calculations are used to determine the mixing enthalpy, cohesive energy, the spin polarised density of states, mechanical properties, and Mulliken’s bond population of the Al13(Fe,M)4 (M = Co, Cu, Cr, Mn, and Ni) phases to better understand how stability, electronic structure, and mechanical characteristics of -Al13Fe4 vary with different alloying elements doping. The Cambridge Serial Total Energy Package (CASTEP) code [23] implements first-principles calculations based on the density functional theory (DFT), with a 3 × 3 × 2 Monkhorst–Pack k-point grid and plane-wave cutoff energy of 500 eV, following a convergence test. The interactions of valence electrons and ionic cores are shown by norm-conserving pseudopotentials (NCPPs). For Al, Fe, Mn, Cr, Co, Ni, and Cu, the valence electrons taken into consideration include 3s23p1, 3d74s1, 3d54s2, 3d54s1, 3d54s1, 3d84s2, and 3d104s1 [24]. The crystal structure is optimized using the Broyden–Fletcher–Goldfarb–Shanno (BFGS) method until the force per atom is lower than 0.02 eV/Å and the total energy change converges to 1 × 10−5 eV. In this study, exchange-correlation energy calculations are performed using the generalized gradient approximation (GGA) within the Perdew, Burke, and Ernzerhof solid approach (PBEsol) scheme [25], which is improved for the surfaces of densely packed solids and the solids themselves. The atomic coordinates and lattice parameters are relaxed along with all of the structures. By deformation of the unit cell, the elastic constant matrix is calculated by making use of the established stress–strain ratios based on generalized Hooke’s law [26]. The Voigt–Reuss–Hill approximation yields the elastic modulus [27].
Thermodynamic calculations are performed using thermo-calc software to explore the precipitate’s solidification sequences in the Al-Mg-Fe alloys based on the TCAL6 database.

3. Results and Discussions

3.1. Determination of Composition and Crystal Structure of Al13Fe4 Phase

Figure 1a shows the calculated solidification sequence and phase fraction of the Al-4.5Mg-1.8Fe alloy in this study. In light of the calculation results, it is evident that θ-Al13Fe4 first precipitates as a primary phase at 664 °C, then eutectic α-Al + θ-Al13Fe4 and Al3Mg2 precipitate at 634 °C and 451 °C, respectively. The microstructure of Al-Mg-Fe alloy has been shown in Figure 1b. It indicates that the primary θ-Al13Fe4 phase manifests an acicular morphology. The mass percentages of Al and Fe in θ-Al13Fe4 are 64.77% and 35.23%, respectively (Figure 1c). The existence of the θ-Al13Fe4 phase is further confirmed by EBSD mapping in Figure 1d, which indicates that the microstructure contains θ-Al13Fe4 (red) and α-Al (green) regions. To identify the crystal structures of the θ-Al13Fe4 accurately, the θ-Al13Fe4 phase in Al-4.56Mg-1.81Fe alloy is examined using TEM. The TEM bright field image of the θ-Al13Fe4 phase has been presented in Figure 1e. The contrast of θ-Al is bright, and it displays a dark strip. Figure 1f displays the high-resolution lattice and atomic picture of θ-Al13Fe4, and Figure 1g inserts the corresponding selected area electron diffraction (SAED) pattern (f). The monoclinic crystal structure of primary -Al13Fe4 with the C2/m space group was discovered using TEM analysis from at least three different zone directions. The lattice parameter of the θ-Al13Fe4 phase is ascertained as a = 15.43 ± 0.01 Å, b = 8.03 ± 0.01 Å, c = 12.44 ± 0.01 Å, presenting a slight deviation from the documented value as a = 15.49 ± 0.01 Å, b = 8.07 ± 0.01 Å, c = 12.47 ± 0.01 Å in the literature [26,27] and a = 14.54 ± 0.01 Å, b = 7.95 ± 0.01 Å, c = 12.42 ± 0.01 Å in the literature [3,16,26]. There is a disparity between the calculation conditions and the experimental conditions of thermoccalc. However, in this paper, we try to ensure that the solidification speed of the alloy is slow as far as possible, and that the experimental conditions are as close to the steady state as possible when preparing the alloy. Although the equilibrium conditions are still not reached, the calculated phase should be consistent with the experimental phase, so it makes sense for DFT calculation. The results of the composition and crystal analysis furnish the necessary modeling data for the first-principles calculations that are to follow. Aluminium and iron each have their own unique set of atom coordinates, which may be found in [26]. The unit cell of θ-Al13Fe4 has a total of 102 atoms, 24 of which are iron, whereas 78 are aluminium [27]. The crystal structure of a 25% atom substitution at Fe sites by M (where M can be any of the elements Co, Cu, Cr, Mn, or Ni) has been established and has been presented θ in Figure 2. The coordinates of the substituted atoms of Fe, as well as the calculated crystal lattice constants and unit cell volume of θ-Al13Fe4, can be found in Table 2.

3.2. Stability and Electronic Structure of Al13(Fe,M)4

The stability of the Al13(Fe,M)4 phase can be estimated via the calculation of the cohesive energy and mixing enthalpy. This energy parameter can be defined in the following equations:
Δ H r A l 13 ( F e , M ) 4 = E t o t A l 13 ( F e , M ) 4 x E b i n A l y E b i n M z E b i n F e x + y + z
E c o h A l 13 ( F e , M ) 4 = E t o t A l 13 ( F e , M ) 4 x E i s o A l y E i s o M z E i s o F e x + y + z
where Δ H r A l 13 ( F e , M ) 4 and E c o h A l 13 ( F e , M ) 4   are the mixing enthalpy and cohesive energy of Al13(Fe,M)4 phases. Etot(Al13(Fe,M)4) is the total energy of Al13(Fe,M)4 per unit cell, and Ebin(Al,Fe,M) is the chemical potential of pure Al, Fe, and M atom. Eiso(Al,Fe,M) is the total energy of a single Al, Fe, and M atom. x, y, and z mean the number of Al, M, and Fe atoms. Al13(Fe,M)4′s cohesive energy and mixing enthalpy findings from calculations are displayed in Table 3. The alloying elements are easier to substitute in the Al13Fe4 phase when the mixing enthalpy is less than zero. The mixing enthalpy for Al13Fe4, Al13(Fe,Co)4, Al13(Fe,Cu)4, Al13(Fe,Cr)4, Al13(Fe,Mn)4, and Al13(Fe,Ni)4 are −0.28, −0.34, −0.39, −0.15, −0.34, and −0.3 eV/atom as shown by the calculated results in Table 3. It shows that all alloy elements can easy to dissolve in Al13(Fe,M)4 phases. The calculated cohesive energy for Al13Fe4, Al13(Fe,Co)4, Al13(Fe,Cu)4, Al13(Fe,Cr)4, Al13(Fe,Mn)4, and Al13(Fe,Ni)4 is −4.96, −5.05, −5.04, −4.92, −4.97, and −4.98 eV/atom, which indicates an improved interatomic bonding force following Co, Cu, Mn, and Ni doping and corroborate well with mixing enthalpy calculation results.
First-principles computation has the benefit of illuminating the physical mechanism underlying the macroscopic physical characteristics of the electronic scale. The partial density of states (PDOS) and total density of states (TDOS) are explored by the computed electronic structures in an attempt to comprehend and elaborate on the physical process of the multialloying impact on the mechanical characteristics of Al13(Fe,M)4 phases, as illustrated in Figure 3. It implies that all Al13(Fe,M)4 phases exhibit metallic character as a result of positive TDOS values at the Fermi level. Fe-3d bands dominate the TDOS in the −3.7 eV to 1.7 eV energy range for all Al13(Fe,M)4 phases. In addition, Al-3p and Fe-3d bands overlap in the −3.7 to 1.7 eV energy range, indicating that these two states have hybridized to produce two bond stats of bonds at various energies. After doping of Co, Cr, and Mn the hybridizations of Al-3p and Co-3d in −2.5 eV to 0 eV energy range, Al-3p and Cr-3d in −2 eV to 0 eV energy range, Al-3p and Mn-3d in −2 eV to 0 eV energy range, contribute to the strong chemical bonding of Al-Co, Al-Cr, and Al-Mn. Al-p bands do not correspond with the Cu and Ni-d bands, which is a sign of a lesser bonding strength than a covalent link, confirming the atomic character of the Cu and Ni doping. Al13(Fe,M)4 phases with 20% Co, Cu, Cr, Mn, and Ni doping are not magnetic, according to the approximately symmetric spin DOS in Figure 3.
The spatial localization of the electron probability density distribution, as represented in the domains assigned to localized bond and lone pair electrons, is graphically depicted by the electron density difference [28,29], the electron density difference function is the difference between the molecular electronic density and the superimposed densities of the constituent non-interacting atoms. The maps of Al13(Fe,M)4 phases on the (001) plane with Co, Cu, Cr, Mn, and Ni doping are shown in Figure 4. The fact that more electrons are localized in a circular pattern around the Co, Cu, Cr, Mn, and Ni atoms than around the Al atoms indicates that these metals have covalent interactions with Co, Cu, Cr, Mn, and Ni.

3.3. Mechanical Properties and Anisotropy

The elastic and ductile–brittle characteristics of θ-Al13Fe4 phases are crucial for the overall performance and applications of Al-Mg-Fe alloys since they are a necessary strengthening phase in Al-Mg-Fe alloys. Their elastic constants have been computed and displayed in Table 4 to examine the impact of Co, Cu, Cr, Mn, and Ni doping on the mechanical characteristics evolution of Al13(Fe,M)4 phases. Born-lattice Huang’s dynamical theory allows us to express the mechanical stability of Al13(Fe,M)4 criteria as:
C11 > 0; C22 > 0; C33 > 0; C44 > 0; C55 > 0; C66 > 0;
C11 + C22 + C33 + 2 (C12 + C13 + C23) > 0;
C11 + 2C12 > 0; (C44C66C462) > 0;
C22 (C33C55C352) + 2C23C25C35C232C35C252C33 > 0;
2 [C15C25 (C33C12C13C23) + C15C35 (C22C13C12C23) + C25C35 (C11C23C12C13)] − [C152 (C11C22C122) + C55 (C11C22C33C232C11C132C22C122C33 + C12C13C23)] > 0
According to Table 4, the calculated elastic constants of these Al13(Fe,M)4 phases meet the aforementioned criteria, indicating that they are mechanically stable at 0 K when doped with Mn, Cr, Co, Ni, and Cu.
The following equations for Al13(Fe,M)4 phases compute the shear and bulk modulus based on the elastic constant:
B V = 1 9 2 C 11 + C 12 + C 23 + C 13 + 22 + C 33  
G V = 1 15 C 11 + C 12 + C 33 C 12 C 13 C 23 + 1 5 C 44 + C 55 + C 66
B R = 1 / S 11 + S 22 + S 33 + 2 S 12 + S 13 + S 23
G R = 1 / 4 S 11 + S 22 + S 33 4 S 12 + S 13 + S 23 + 3 S 44 + S 55 + S 66
where Cij is the elastic constant and Sij = 1/Cij, GV (GR) and BV (BR) are the shear modulus and bulk modulus computed from Voigt (Reuss) model. Therefore, the calculation of their average value gives the final values of B and G [30]:
B = B V + B R / 2
G = G V + G R / 2
The Young’s modulus (E) and Poisson’s ratio (σ) of Al13(Fe,M)4 can be obtained as follows:
E = 9 B G / 3 B + G
σ = 3 B 2 G / 6 B + 2 G
The intrinsic brittleness of Fe-containing intermetallics substantially hinders the application of Al-Mg-Fe alloys. Due to the lack of experimental values, we have compared the calculated values with other work, which shows a good agreement as illustrated in Table 5. It is, hence, indispensable to investigate the enhancement of its ductility via doping of alloying elements. The mechanical properties evolution of Al13(Fe,M)4 phases with Co, Cu, Cr, Mn, and Ni and doping have been summarized in Table 4 and Figure 5. The critical values of Poisson’s ratio (0.26) and Pugh ratio B/G (1.75) have been broadly employed as indicators representing the ductility of materials. A larger Poisson’s ratio (σ) signifies the higher softness of the material [31]. The σ and B/G values for Al13Fe4, Al13(Fe,Cu)4, and Al13(Fe,Ni)4 are 0.273, 0.317, 0.269, and 1.87, 2.41, 1.83, respectively, which means a better ductility of these phases.
The capacity of Al13(Fe,M)4 phases to stop cracks from spreading is characterized by their fracture toughness (KIC), which is crucial for the service safety of Al-Mg-Fe alloys. Ogata and Niu have developed a model for determining fracture toughness that can be applied to several materials by examining the relationship between the elastic characteristics of materials and their fracture toughness. During the deformation of a material, bonds break and reform resulting in the displacement of atoms and slipping of atomic planes, and materials with high fracture toughness usually exhibit high ductility and yield at high critical strain. Therefore, B/G is in positive correlation with fracture toughness KIC [32]. This paper uses the following, reasonably straightforward calculation for fracture toughness [33,34,35,36]:
K IC = V 0 1 / 6 · G · B / G 1 / 2  
where G and B, respectively, denote the shear and bulk moduli (in MPa) of Al13(Fe,M)4, V0 is the volume per atom (in m3), and the unit of KIC is MPa m1/2. By comparison of the KIC in Table 5, it was found that Co, Cr, and Mn are prospective candidates for the fracture toughness optimization of Al13Fe4.
The mechanical modulus value and variation of Al13(Fe,M)4 phases with the doping of different alloying elements are listed in Table 5 and Figure 6. The calculated Young’s modulus value of Al13Fe4 is 158.41 GPa. With the doping of Co, Cr, and Mn, it is increased to 181.25, 180.82, and 187.48 GPa. However, for the doping of Cu and Ni, the calculated Young’s modulus value of Al13(Fe,M)4 decreased to 129.58 and 154.43GPa, respectively. The shear and bulk moduli both exhibit the same fluctuation trend. Young’s modulus can reveal the degree of atom bonding in crystals. In general, the strength of a chemical bond increases with increasing Young’s modulus values. The mechanical modulus evolution can therefore be attributed to the variation of chemical bonding strength in Al13(Fe,M)4, which is also validated by the TDOS in Figure 4. It can be concluded that the chemical bonding strength of Al-M increases in the sequence of Al-Cu, Al-Ni, Al-Fe, Al-Cr, Al-Co, and Al-Mn, which can be appropriately elaborated using the mean bond population variation in Figure 6.
An essential index for the mechanical assessment of Al13Fe4 is its intrinsic hardness. In this study, two alternative models—model Chen’s (HVC) and Tian’s model (HVT)—are used to predict the hardness evolution of Al13(Fe,M)4 with Co, Cu, Cr, Mn, and Ni doping [39].
H V C = 2 ( k 2 G ) 0.585 3
H V T = 0.92 k 1.137 G 0.708
The Pugh ratio, k, is defined as k = B/G. Table 5 provides a summary of the computed outcomes. As shown in Figure 6, the computed results of the elastic moduli, B/G, and HV are deducted from the corresponding Al13Fe4 values to more clearly illustrate the differences in the mechanical characteristics of Al13(Fe,M)4 (i.e., Al13Fe4 is taken as the reference state). The altered mechanical characteristics of Al13(Fe,M)4 brought on by the doping of alloying elements are then discovered. The G/B and values fluctuation show that Cu doping can increase Al13Fe4′s ductility while lowering its elastic modulus and hardness. Co, Cr, and Mn can increase the elastic modulus and hardness prominently.
The mean bond population and bond length variation of Al-Fe and Al-M in Al13(Fe,M)4 phases with Co, Cu, Cr, Mn, and Ni doping are depicted in Figure 6. Mulliken atomic populations are largely based on first-order electron density functions within a linear combination of atomic orbitals–molecular orbital (LCAO-MO) theory [39]. Mulliken population analysis assesses the distribution of electrons in several fractional means among the different parts of atomic bonds. In addition, the overlap population abides by a good relationship with the covalency of bonding and bond strength. More description is given in [21,22]. Mean bong population is an analysis of the electronic occupancy state in atomic orbits, which can be calculated as follow:
n m A B = k W k u o n A v o n B P u v k S u v k
where W(k) is the weight, k is the wave vector, u and v are the electron orbitals, P(k) is the density matrix of the corresponding electron orbitals, and S(k) is the overlap matrix. The average bond length and the average number of bonds can be obtained by the following equation:
L ¯ ( A B ) = i L i N i i N i
n ¯ ( A B ) = i n i A B N i i N i
where Ni is the total number of different bonds in the unit cell, Li is the bond length of different bonds, and ni is the population of different bonds. The bond length and population of the Al-Fe and Al-M bonds in Al13Fe4 were changed by the alloying elements. Larger and more positive population values typically suggest the existence of stronger covalent bonds. For every Fe atom that is replaced by an alloying atom in Al13Fe4, one Al-M bond is replaced by a pair of original Al-Fe bonds. The bond number and bond length of the Al-M bonds in Al13Fe4 were changed by the alloying metals. Al13(Fe,M)4 has all Al-M bonds, although the chemical linkages Al-Co, Al-Cr, and Al-Mn have the largest mean bond populations (Figure 6), indicating a stronger bonding strength. This precisely explains the improved modulus value of Al13(Fe,Co)4, Al13(Fe,Cr)4, and Al13(Fe,Mn)4 among these Al13(Fe,M)4 phases. For Al13(Fe,Cu)4 and Al13(Fe,Ni)4, the smaller mean bond population of Al-Cu and Al-Ni indicates their weaker bonding strength, leading to the smaller mechanical modulus and hardness.
Since crack nucleation and propagation are directly related to elastic anisotropy, Al13Fe4′s mechanical characteristics anisotropy is also crucial for the performance of Al-Mg-Fe alloys. The Young’s modulus, hardness, and Poisson ratio surface contours of -Al13Fe4 with Co, Cu, Cr, Mn, and Ni doping are depicted in Figure 7, Figure 8 and Figure 9, respectively, to illustrate the elastic anisotropy of Al13(Fe,M)4. The mechanical surface contours of isotropic materials are a perfect sphere. Al13Fe4 exhibits high Young’s modulus anisotropy, according to Figure 7a, since the surface contours depart from a perfect sphere. Young’s modulus values along [001] directions are higher than other directions for all Al13(Fe,M)4 phases. The modulus anisotropy of Al13(Fe,M)4 phases can be predicted using the computed Cij in Table 4. C11 and C33 represent a crystal’s capacity to withstand axial strain in the [100] and [001] axes, respectively [17,21,35,36,37,38]. The modulus of Al13(Fe,M)4 phases along the [100] direction is smaller than that in the [001] directions, as indicated by C11 being smaller than C33. The results are confirmed by modulus surface contours in Figure 7. Doping Al13Fe4 with elements such as Co, Cu, Cr, and Ni results in a strengthening of the material’s elastic anisotropy (Figure 7b–d,f). Mn is an interesting candidate for use as an alloying element because of its potential to reduce the elastic anisotropy of Al13Fe4 (see Figure 7e). This, in turn, would be beneficial to the mechanical characteristics of Al-Mg-Fe alloys (Figure 7f). As can be seen in Figure 8, the bulk modulus of these Al13(Fe,M)4 phases exhibit a clear anisotropy in their behavior. The value of all Al13(Fe,M)4 phases is greatest along the direction indicated by the notation [011]. Doping with Co and Cu led to a reduction in the elastic anisotropy (Figure 8b,c), whereas doping with Ni resulted in an extensive strengthening of the elastic anisotropy (Figure 8f).
The Poisson ratio surface contours of Al13(Fe,M)4 with various alloying elements doped are plotted in Figure 9 to more clearly illustrate the characteristics of mechanical anisotropy. It is obvious that Al13(Fe,Ni)4 shows the strongest Possion ratio anisotropy, and the descending order of Poisson ratio anisotropy for Al13(Fe,M)4 phases is Al13(Fe,Ni)4 > Al13(Fe,Mn)4 > Al13Fe4 > Al13(Fe,Cr)4 > Al13(Fe,Co)4 > Al13(Fe,Cu)4.

4. Conclusions

In conclusion, using first-principles calculations based on DFT, the impact of Co, Cr, Cu, Mn, and Ni doping on the mechanical characteristics and electronic structure of Al13Fe4 has been comprehensively studied. The following are the primary conclusions:
(1)
The substitution of Fe by M (M = Co, Cr, Cu, Mn, and Ni) in Al13Fe4 is energetically favourable at 0 K due to the negative mixing enthalpy and cohesive energy.
(2)
With the doping of Co, Cr, and Mn, the Young’s modulus values are increased to 181.25, 180.82, and 187.48 GPa, the hardness values are increased to 10.9, 10.95, and 11.12 GPa, owing to the formation of stronger Al-Co, Al-Cr, and Al-Mn chemical bonds, the existence of which is validated by electronic structure calculation conducted in this study.
(3)
Co, Cr, and Mn are potential candidate elements for the mechanical optimization of Al13Fe4 for the simultaneous improvement of mechanical modulus, hardness, and fracture toughness (KIC from 1.34 to 1.46, 1.45, and 1.49).

Author Contributions

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

Funding

This research received no external funding.

Data Availability Statement

All data, models, and code generated or used during the study appear in the submitted article.

Conflicts of Interest

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

References

  1. Stefániay, V.; Griger, A.; Turmezey, T. Intermetallic phases in the aluminium-side corner of the AlFeSi-alloy system. J. Mater. Sci. 1987, 22, 539–546. [Google Scholar] [CrossRef]
  2. Rømming, C.; Hansen, V.; Gjønnes, J. Crystal structure of β-Al4.5FeSi. Acta Crystallogr. Sect. B Struct. Sci. 1994, 50, 307–312. [Google Scholar] [CrossRef]
  3. Que, Z.; Mendis, C.L. Heterogeneous nucleation and phase transformation of Fe-rich intermetallic compounds in Al–Mg–Si alloys. J. Alloys Compd. 2020, 836, 155515. [Google Scholar] [CrossRef]
  4. Zhu, X.; Blake, P.; Dou, K.; Ji, S. Strengthening die-cast Al-Mg and Al-Mg-Mn alloys with Fe as a beneficial element. Mater. Sci. Eng. A 2018, 732, 240–250. [Google Scholar] [CrossRef]
  5. Cai, B.; Kao, A.; Lee, P.; Boller, E.; Basevi, H.; Phillion, A.; Leonardis, A.; Pericleous, K. Growth of β intermetallic in an Al-Cu-Si alloy during directional solidification via machine learned 4D quantification. Scr. Mater. 2019, 165, 29–33. [Google Scholar] [CrossRef]
  6. Puncreobutr, C.; Lee, P.; Kareh, K.; Connolley, T.; Fife, J.; Phillion, A. Influence of Fe-rich intermetallics on solidification defects in Al–Si–Cu alloys. Acta Mater. 2014, 68, 42–51. [Google Scholar] [CrossRef]
  7. Terzi, S.; Taylor, J.A.; Cho, Y.; Salvo, L.; Suéry, M.; Boller, E.; Dahle, A.K. Nucleation and growth dynamics of the α-Al/β-Al5FeSi eutectic in a complex Al-Si-Cu-Fe alloy. In Proceeding of the 12th International Conference on Alu-minium Alloys, Yokohama, Japan, 5–9 September 2010; pp. 1273–1278. [Google Scholar]
  8. Ferraro, S.; Timelli, G. Influence of Sludge Particles on the Tensile Properties of Die-Cast Secondary Aluminum Alloys. Met. Mater. Trans. A 2014, 46, 1022–1034. [Google Scholar] [CrossRef]
  9. Gao, T.; Hu, K.; Wang, L.; Zhang, B.; Liu, X. Morphological evolution and strengthening behavior of α-Al(Fe,Mn)Si in Al–6Si–2Fe–xMn alloys. Results Phys. 2017, 7, 1051–1054. [Google Scholar] [CrossRef]
  10. Becker, H.; Bergh, T.; Vullum, P.; Leineweber, A.; Li, Y. Effect of Mn and cooling rates on α-, β- and δ-Al–Fe–Si intermetallic phase formation in a secondary Al–Si alloy. Materialia 2018, 5, 100198. [Google Scholar] [CrossRef]
  11. Zhenming, X.; Tianxiao, L.; Yaohe, Z. Elimination of Fe in Al-Si cast alloy scrap by electromagnetic filtration. J. Mater. Sci. 2003, 38, 4557–4565. [Google Scholar] [CrossRef]
  12. Rajabi, M.; Vahidi, M.; Simchi, A. Microstructural evolution of Al-20Si-5Fe alloy during rapid solidification and hot consoli-dation. Rare Metals 2009, 28, 639–645. [Google Scholar] [CrossRef]
  13. Seifeddine, S.; Johansson, S.; Svensson, I.L. The influence of cooling rate and manganese content on the β-Al5FeSi phase formation and mechanical properties of Al–Si-based alloys. Mater. Sci. Eng. A 2008, 490, 385–390. [Google Scholar] [CrossRef]
  14. Ferreira, T.; Vidilli, A.L.; Zepon, G.; Kiminami, C.S.; Botta, W.J. Bolfarini, Strong and ductile recycled Al-7Si-3Cu-1Fe alloy: Controlling the morphology of quasicrystal approximant α-phase by Mn and V addition. J. Alloys Compd. 2021, 888, 161508. [Google Scholar] [CrossRef]
  15. Zhao, Y.L.; Song, D.F.; Wang, H.L.; Jia, Y.W.; Lin, B.; Tang, Y.; Tang, Y.; Shu, D.; Sun, Z.Z.; Fu, Y.N.; et al. Revealing the influence of Fe on Fe-rich phases formation and mechanical properties of cast Al-Mg-Mn-Fe alloys. J. Alloys Compd. 2022, 901, 163666. [Google Scholar] [CrossRef]
  16. Fang, C.M.; Que, Z.P.; Dinsdale, A.; Fan, Z. Si solution in θ-Al13Fe4 from first-principles. Intermetallics 2020, 126, 106939. [Google Scholar] [CrossRef]
  17. Chong, X.; Palma, J.P.S.; Wang, Y.; Shang, S.-L.; Drymiotis, F.; Ravi, V.A.; Star, K.E.; Fleurial, J.-P.; Liu, Z.-K. Thermodynamic properties of the Yb-Sb system predicted from first-principles calculations. Acta Mater. 2021, 217, 117169. [Google Scholar] [CrossRef]
  18. Popcevic, P.; Smontara, A.; Ivkov, J.; Wencka, M.; Komelj, M.; Vrtnik, S.; Bobnar, M.; Jaglicic, Z.; Bauer, B.; Gille, P.; et al. Anisotropic physical properties of the Al13Fe4 complex intermetallic and its ternary derivative Al13(Fe,Ni)4. Phys. Rev. B 2010, 81, 184203. [Google Scholar] [CrossRef]
  19. Zhang, X.; Wang, D.; Zhou, Y.; Chong, X.; Li, X.; Zhang, H.; Nagaumi, H. Exploring crystal structures, stability and mechanical properties of Fe, Mn-containing intermetallics in Al-Si Alloy by experiments and first-principles calculations. J. Alloys Compd. 2021, 876, 160022. [Google Scholar] [CrossRef]
  20. Chen, Y.; Yao, Z.J.; Zhang, P.Z.; Wei, D.B.; Luo, X.X.; Xue, F. Effect of Alloying Elements V, Cr and Ni on the Electronic Structure and Mechanical Properties of FeAl from First-Principles Calculation. Adv. Mater. Res. 2014, 887–888, 378–383. [Google Scholar] [CrossRef]
  21. Chong, X.; Hu, M.; Wu, P.; Shan, Q.; Jiang, Y.H.; Li, Z.L.; Feng, J. Tailoring the anisotropic mechanical properties of hexagonal M7X3 (M=Fe, Cr, W, Mo; X=C, B) by multialloying. Acta Mater. 2019, 169, 193–208. [Google Scholar] [CrossRef]
  22. Chong, X.; Guan, P.; Hu, M.; Jiang, Y.H.; Li, Z.L.; Feng, J. Exploring accurate structure, composition and thermophysical properties of η carbides in 17.90 wt% W-4.15 wt% Cr-1.10 wt% V-0.69 wt% C steel. Scr. Mater. 2018, 154, 149–153. [Google Scholar] [CrossRef]
  23. Wang, G.; Jiang, Y.; Li, Z.; Chong, X.; Feng, J. Balance between strength and ductility of dilute Fe2B by high-throughput first-principles calculations. Ceram. Int. 2020, 47, 4758–4768. [Google Scholar] [CrossRef]
  24. Segall, M.D.; Lindan, P.J.D.; Probert, M.J.; Pickard, C.J.; Hasnip, P.J.; Clark, S.J.; Payne, M.C. First-principles simulation: Ideas, illustrations and the CASTEP code. J. Phys. Condens. Matter 2002, 14, 2717–2744. [Google Scholar] [CrossRef]
  25. Wei, X.; Chen, Z.; Zhong, J.; Wang, L.; Yang, W.; Wang, Y. Effect of alloying elements on mechanical, electronic and magnetic properties of Fe2B by first-principles investigations. Comput. Mater. Sci. 2018, 147, 322–330. [Google Scholar] [CrossRef]
  26. Perdew, J.P.; Ruzsinszky, A.; Csonka, G.I.; Vydrov, O.A.; Scuseria, G.E.; Constantin, L.A.; Zhou, X.; Burke, K. Restoring the Density-Gradient Expansion for Exchange in Solids and Surfaces. Phys. Rev. Lett. 2008, 100, 136406. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Shang, S.; Wang, Y.; Liu, Z.-K. First-principles elastic constants of α- and θ-Al2O3. Appl. Phys. Lett. 2007, 90, 101909. [Google Scholar] [CrossRef]
  28. Wu, Z.-J.; Zhao, E.-J.; Xiang, H.-P.; Hao, X.-F.; Liu, X.-J.; Meng, J. Crystal structures and elastic properties of superhardIrN2andIrN3from first principles. Phys. Rev. B 2007, 76, 054115. [Google Scholar] [CrossRef]
  29. Gueneau, C.; Servant, C.; D’Yvoire, F.; Rodier, N. FeAl3Si2. Acta Crystallogr. Sect. C Cryst. Struct. Commun. 1995, 51, 177–179. [Google Scholar] [CrossRef]
  30. Tsuchiya, T.; Yamanaka, T.; Matsui, M. Molecular dynamics study of pressure-induced transformation of quartz-type GeO2. Phys. Chem. Miner. 2000, 27, 149–155. [Google Scholar] [CrossRef]
  31. Chong, X.; Jiang, Y.; Zhou, R.; Feng, J. Elastic properties and electronic structures of CrxBy as superhard compounds. J. Alloys Compd. 2014, 610, 684–694. [Google Scholar] [CrossRef]
  32. Niu, H.; Niu, S.; Oganov, A.R. Simple and accurate model of fracture toughness of solids. J. Appl. Phys. 2019, 125, 065105. [Google Scholar] [CrossRef] [Green Version]
  33. Mazhnik, E.; Oganov, A.R. A model of hardness and fracture toughness of solids. J. Appl. Phys. 2019, 126, 125109. [Google Scholar] [CrossRef]
  34. Li, Y.; Liu, Y.; Yang, J. First principle calculations and mechanical properties of the intermetallic compounds in a laser welded steel/aluminum joint. Opt. Laser Technol. 2020, 122, 105875. [Google Scholar] [CrossRef]
  35. He, L.; Li, Z.; Zhang, J.; Peng, F.; Zhao, S.; Chen, H.; Yan, H.; Yang, T.; Chen, S.; Liu, B.; et al. Softening Al13Fe4 intermetallic compound through Fe-site multi-principal-element doping. Scr. Mater. 2022, 218, 114811. [Google Scholar] [CrossRef]
  36. Chong, X.; Jiang, Y.; Hu, M.; Feng, J. Elaborating the phases and mechanical properties of multiphase alloy: Experimental two-dimensional mapping combined with theoretical calculations. Mater. Charact. 2017, 134, 347–353. [Google Scholar] [CrossRef]
  37. Chen, X.-Q.; Niu, H.; Li, D.; Li, Y. Modeling hardness of polycrystalline materials and bulk metallic glasses. Intermetallics 2011, 19, 1275–1281. [Google Scholar] [CrossRef] [Green Version]
  38. Tian, Y.; Xu, B.; Zhao, Z. Microscopic theory of hardness and design of novel superhard crystals. Int. J. Refract. Met. Hard Mater. 2012, 33, 93–106. [Google Scholar] [CrossRef]
  39. Chong, X.; Jiang, Y.; Zhou, R.; Feng, J. Multialloying effect on thermophysical properties of Cr7 C3 -type carbides. J. Am. Ceram. Soc. 2017, 100, 1588–1597. [Google Scholar] [CrossRef]
Figure 1. (a) Calculated solidification sequence of Al-4.5Mg-1.8Fe alloy, (b) microstructures and EDS map of Al alloy, (c) composition of Al13Fe4, (d) EBSD maps of A1 alloys, (e) TEM bright field photograph of Al13Fe4 phase; (f) HRTEM photograph, and (g) the corresponding SAED pattern of Al13Fe4 phase.
Figure 1. (a) Calculated solidification sequence of Al-4.5Mg-1.8Fe alloy, (b) microstructures and EDS map of Al alloy, (c) composition of Al13Fe4, (d) EBSD maps of A1 alloys, (e) TEM bright field photograph of Al13Fe4 phase; (f) HRTEM photograph, and (g) the corresponding SAED pattern of Al13Fe4 phase.
Metals 12 01999 g001
Figure 2. Crystal structure projection of Al13(Fe,M)4 phases with different alloying elements doping. (a) θ-Al13Fe4, (b) Al13(Fe,Co)4, (c) Al13(Fe,Cu)4, (d) Al13(Fe,Cr)4, (e) Al13(Fe,Mn)4, (f) Al13(Fe,Ni)4.
Figure 2. Crystal structure projection of Al13(Fe,M)4 phases with different alloying elements doping. (a) θ-Al13Fe4, (b) Al13(Fe,Co)4, (c) Al13(Fe,Cu)4, (d) Al13(Fe,Cr)4, (e) Al13(Fe,Mn)4, (f) Al13(Fe,Ni)4.
Metals 12 01999 g002
Figure 3. The spin-polarized total density of states (TDOS) and partial density of states (PDOS) of Al13(Fe,M)4 phases with Co, Cu, Cr, Mn, and Ni doping. Dashed lines represent the Fermi level. (a) Al13Fe4, (b) Al13(Fe,Co)4, (c) Al13(Fe,Cu)4, (d) Al13(Fe,Cr)4, (e) Al13(Fe,Mn)4, (f) Al13(Fe,Ni)4.
Figure 3. The spin-polarized total density of states (TDOS) and partial density of states (PDOS) of Al13(Fe,M)4 phases with Co, Cu, Cr, Mn, and Ni doping. Dashed lines represent the Fermi level. (a) Al13Fe4, (b) Al13(Fe,Co)4, (c) Al13(Fe,Cu)4, (d) Al13(Fe,Cr)4, (e) Al13(Fe,Mn)4, (f) Al13(Fe,Ni)4.
Metals 12 01999 g003
Figure 4. The electron density difference distribution of Al13(Fe,M)4 phases with different elements doping on the plane (010). (a) Al13Fe4, (b) Al13(Fe,Co)4, (c) Al13(Fe,Cu)4, (d) Al13(Fe,Cr)4, (e) Al13(Fe,Mn)4, (f) Al13(Fe,Ni)4.
Figure 4. The electron density difference distribution of Al13(Fe,M)4 phases with different elements doping on the plane (010). (a) Al13Fe4, (b) Al13(Fe,Co)4, (c) Al13(Fe,Cu)4, (d) Al13(Fe,Cr)4, (e) Al13(Fe,Mn)4, (f) Al13(Fe,Ni)4.
Metals 12 01999 g004
Figure 5. The multialloying effect on the mechanical properties of Al13(Fe,M)4 phases with different alloying elements doping; (a) the bulk, shear, and Young’s modulus; (b) the B/G; (c) the Poisson’s ratio (σ); (d) the intrinsic hardness based on Tian and Chen model. It must be noted that Al13Fe4 is set as the reference and all relative values are taken with respect to those of the Al13Fe4 phase. Additionally, the negative values refer to the decrease in the properties of interest upon the addition of alloying elements to Al13Fe4.
Figure 5. The multialloying effect on the mechanical properties of Al13(Fe,M)4 phases with different alloying elements doping; (a) the bulk, shear, and Young’s modulus; (b) the B/G; (c) the Poisson’s ratio (σ); (d) the intrinsic hardness based on Tian and Chen model. It must be noted that Al13Fe4 is set as the reference and all relative values are taken with respect to those of the Al13Fe4 phase. Additionally, the negative values refer to the decrease in the properties of interest upon the addition of alloying elements to Al13Fe4.
Metals 12 01999 g005
Figure 6. The average bond length and average bond population of Al-Fe bonds (a) and Al-M bonds (b) in Al13(Fe,M)4.
Figure 6. The average bond length and average bond population of Al-Fe bonds (a) and Al-M bonds (b) in Al13(Fe,M)4.
Metals 12 01999 g006
Figure 7. Young’s modulus 3D surface of Al13(Fe,M)4 phases with different elements doping (unit: GPa). (a) Al13Fe4, (b) Al13(Fe,Co)4, (c) Al13(Fe,Cu)4, (d) Al13(Fe,Cr)4, (e) Al13(Fe,Mn)4, (f) Al13(Fe,Ni)4.
Figure 7. Young’s modulus 3D surface of Al13(Fe,M)4 phases with different elements doping (unit: GPa). (a) Al13Fe4, (b) Al13(Fe,Co)4, (c) Al13(Fe,Cu)4, (d) Al13(Fe,Cr)4, (e) Al13(Fe,Mn)4, (f) Al13(Fe,Ni)4.
Metals 12 01999 g007
Figure 8. Bulk modulus 3D surface of Al13(Fe,M)4 phases with different elements doping (unit: GPa). (a) Al13Fe4, (b) Al13(Fe,Co)4, (c) Al13(Fe,Cu)4, (d) Al13(Fe,Cr)4, (e) Al13(Fe,Mn)4, (f) Al13(Fe,Ni)4.
Figure 8. Bulk modulus 3D surface of Al13(Fe,M)4 phases with different elements doping (unit: GPa). (a) Al13Fe4, (b) Al13(Fe,Co)4, (c) Al13(Fe,Cu)4, (d) Al13(Fe,Cr)4, (e) Al13(Fe,Mn)4, (f) Al13(Fe,Ni)4.
Metals 12 01999 g008
Figure 9. Poisson ratio 3D surface of Al13(Fe,M)4 phases with different elements doping. (a) Al13Fe4, (b) Al13(Fe,Co)4, (c) Al13(Fe,Cr)4, (d) Al13(Fe,Cu)4, (e) Al13(Fe,Mn)4, (f) Al13(Fe,Ni)4.
Figure 9. Poisson ratio 3D surface of Al13(Fe,M)4 phases with different elements doping. (a) Al13Fe4, (b) Al13(Fe,Co)4, (c) Al13(Fe,Cr)4, (d) Al13(Fe,Cu)4, (e) Al13(Fe,Mn)4, (f) Al13(Fe,Ni)4.
Metals 12 01999 g009
Table 1. The chemical compositions of alloys investigated in the experiment (wt.%).
Table 1. The chemical compositions of alloys investigated in the experiment (wt.%).
MgFeTiCuSiMnAl
4.561.81<0.05<0.03<0.02<0.03Bal
Table 2. Structural information, lattice parameters (Å), cell volume (Å3), and formation enthalpy (Ef) (at 0 K) for M occupying Fe sites in the θ-Al13Fe4 phase.
Table 2. Structural information, lattice parameters (Å), cell volume (Å3), and formation enthalpy (Ef) (at 0 K) for M occupying Fe sites in the θ-Al13Fe4 phase.
Al13(Fe,M)4Atom Coordinates of MLattice Constants (Å)
abcV (Å3)
Al13Fe4M1 (0.098, 0.5, 0.38)
M2 (0.414, 0.5, 0.62)
M3 (0.59, 0.5, 0.99)
M4 (0.43, 0.5, 0.98)
15.438.0312.441542.5α = 89.9°
β = 107.7°
γ = 90.1°
Al13(Fe,Co)415.418.0612.431543.7
Al13(Fe,Cu)415.428.1512.461566
Al13(Fe,Cr)415.648.0712.571586.6
Al13(Fe,Mn)415.58.0412.51558.2
Al13(Fe,Ni)415.428.0812.421547.6
Table 3. The calculated mixing enthalpy and cohesive energy (eV/atom) of Al13(Fe,M)4 phases with Co, Cu, Cr, Mn, and Ni doping.
Table 3. The calculated mixing enthalpy and cohesive energy (eV/atom) of Al13(Fe,M)4 phases with Co, Cu, Cr, Mn, and Ni doping.
SpeciesCohesive Energy (eV/atom)Mixing Enthalpy (eV/atom)
Al13Fe4−4.69−0.28
Al13(Fe,Co)4−5.05−0.34
Al13(Fe,Cu)4−5.04−0.39
Al13(Fe,Cr)4−4.92−0.15
Al13(Fe,Mn)4−4.97−0.34
Al13(Fe,Ni)4−4.98−0.3
Table 4. Estimated elastic constants (in GPa) of Al13(Fe,M)4 phases with different elements doping.
Table 4. Estimated elastic constants (in GPa) of Al13(Fe,M)4 phases with different elements doping.
SpeciesCij
C11C22C33C44C55C66C12C13C23C15C25C35C46
Al13Fe4225.7219.4235.261.5272.670.683.268.339.8−1.74−2.95.8−2.1
Al13(Fe,Co)4222.1217.4236.164.675.168.282.170.044.0−0.51−3.183.6−4.3
Al13(Fe,Cu)4205.9174.2224.850.468.859.5195.466.342.6−0.42−1.712.2−1.3
Al13(Fe,Cr)4218.7213.2235.657.950.252.581.067.940.9−1.7−2.43.54.7
Al13(Fe,Mn)4236.6214.7240.565.478.676.082.270.646.1−3.7−1.5−4.1−2.0
Al13(Fe,Ni)4221.3203.9239.618.474.040.988.867.945.6−1.75−2.533.15.2
Table 5. Bulk modulus, shear modulus, Young’s modulus, Poisson’s ratio, B/G, and intrinsic hardness of Al13(Fe,M)4 phases with different elements doping, and the unit is GPa.
Table 5. Bulk modulus, shear modulus, Young’s modulus, Poisson’s ratio, B/G, and intrinsic hardness of Al13(Fe,M)4 phases with different elements doping, and the unit is GPa.
SpeciesBGEσB/GHVCHVTKIC (MPa·m1/2)
Al13Fe4116.1162.24158.410.2731.877.818.441.34
Cal. Al13Fe4 1123761890.2431.618---
Cal. Al13Fe4 2121.88686.942210.7230.2111.40---
Al13(Fe,Co)4118.4372.8181.250.2451.6310.911.011.46
Al13(Fe,Cu)4118.4249.17129.580.3172.413.985.341.20
Al13(Fe,Cr)4117.7772.67180.820.2441.6210.9511.051.45
Al13(Fe,Mn)4120.6474.56187.480.2361.6111.1211.231.49
Al13(Fe,Ni)4111.6860.87154.530.2691.837.888.461.30
1—Calculated values of Yulong Li et al. [37]. 2—Calculated values of Liu He et al. [38].
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, Q.; Zhang, H.; Jiang, P.; Lv, Y. The Alloying Strategy to Tailor the Mechanical Properties of θ-Al13Fe4 Phase in Al-Mg-Fe Alloy by First-Principles Calculations. Metals 2022, 12, 1999. https://doi.org/10.3390/met12121999

AMA Style

Liu Q, Zhang H, Jiang P, Lv Y. The Alloying Strategy to Tailor the Mechanical Properties of θ-Al13Fe4 Phase in Al-Mg-Fe Alloy by First-Principles Calculations. Metals. 2022; 12(12):1999. https://doi.org/10.3390/met12121999

Chicago/Turabian Style

Liu, Qianli, Hao Zhang, Peng Jiang, and Yifan Lv. 2022. "The Alloying Strategy to Tailor the Mechanical Properties of θ-Al13Fe4 Phase in Al-Mg-Fe Alloy by First-Principles Calculations" Metals 12, no. 12: 1999. https://doi.org/10.3390/met12121999

APA Style

Liu, Q., Zhang, H., Jiang, P., & Lv, Y. (2022). The Alloying Strategy to Tailor the Mechanical Properties of θ-Al13Fe4 Phase in Al-Mg-Fe Alloy by First-Principles Calculations. Metals, 12(12), 1999. https://doi.org/10.3390/met12121999

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