2022 Volume 63 Issue 10 Pages 1359-1368
Magnesium, the lightest structural metal, needs to improve its strength and formability at room temperature for commercialization. Manganese is an alloying element that can improve the strength of magnesium through forming intermetallic compounds. However, its effect on the formability of magnesium has not yet been clarified. To identify the effect, an interatomic potential for the Mg–Mn binary system has been developed on the basis of the second nearest-neighbor modified embedded-atom method formalism. The Mg–Mn potential reproduces the structural, elastic, and thermodynamic properties of the compound and solution phases of its associated alloy system, consistent with experimental data and higher-level calculations. The applicability of the developed potential is demonstrated by calculating the generalized stacking fault energy for various slip systems of the Mg–Mn alloy.
Fig. 4 Calculated GSFE (mJ/m2) curves of pure Mg and Mg–Mn, models on the (a) basal plane along the $\langle 10\bar{1}0\rangle $ direction, (b) basal plane along the $\langle 11\bar{2}0\rangle $ direction, (c) prismatic plane along the $\langle 11\bar{2}0\rangle $ direction, (d) pyramidal I plane along the $\langle 11\bar{2}0\rangle $ direction, (e) pyramidal I plane along the $\langle 11\bar{2}3\rangle $ direction, and (f) pyramidal II plane along the $\langle 11\bar{2}3\rangle $ direction, using the present 2NN MEAM potential.
Light-weight, high-strength, and high-formability at room temperature (RT) are the main keywords in recent material development, especially in the automotive industry.1,2) Magnesium (Mg), the lightest structural metal, satisfies the first keyword well.3) Despite this, Mg has lower strength and formability at RT than steels or aluminum (Al) alloys.4) Hence, both these properties need to be improved for its commercialization.
A general method to improve the strength of Mg is precipitation hardening.5) Precipitates can be formed by adding alloying elements that form intermetallic compounds with Mg or other frequently added alloying elements, such as Al and zinc (Zn). Among the alloying elements frequently added to Mg, manganese (Mn) satisfies the above conditions. Mn does not form any intermetallic compounds with Mg,6) but forms seven and five intermetallic compounds with Al7) and Zn,8) respectively. These Mn-containing intermetallic compounds of various compositions and structures could improve the strength of Mg by forming many precipitates.
Despite their strengthening abilities, the effect of Mn on the RT formability of Mg is still unclear. It has been reported that this effect of Mn varies depending on the deformation process; an extruded Mg–1 mass% Mn alloy exhibits higher ductility and formability at RT than pure Mg,9) but a hot rolled Mg–1 mass% Mn alloy shows lower ductility at RT than pure Mg.10) Because of the different experimental results, a more fundamental study is necessary to reveal the effect of Mn on Mg formability.
Previous studies have reported that the Mg formability depends on the slip, twin, and recrystallization behaviors.5) These three behaviors are fundamentally related to dislocations that cause a slip, contribute to nucleation and growth of twins,11,12) and nucleate subgrains during recrystallization.13) Hence, we need to investigate the effect of Mn on the dislocation behavior of Mg to further understand its fundamental effect on the Mg formability. The dislocation behavior is a dynamic material phenomenon at an atomic scale that changes with time. Therefore, its observation can be performed using tools that allow a continuous nanoscale analysis of the samples, such as an in situ transmission electron microscope (TEM), a first-principles calculation, and a large-scale atomistic simulation. The in situ TEM can observe the motion of the dislocations in a sample, but it is time- and cost-consuming to observe all kinds of Mg dislocations, such as basal, prismatic, pyramidal I, and pyramidal II dislocations. The first-principles calculation can analyze a basal dislocation with a simple core structure, but analyzing non-basal dislocations with complex core structures is a challenge. The large-scale atomistic simulation is comparatively more appropriate for investigating those various Mg dislocations because it can analyze the movement of millions of atoms at an atomic scale. For this reason, the large-scale atomistic simulation method was chosen to investigate the dislocation behavior of the Mg–Mn alloys.
The large-scale atomistic simulation can be conducted based on (semi-)empirical interatomic potentials describing relevant alloy systems. This means that an interatomic potential for the Mg–Mn binary system is required to perform the dislocation investigation. However, this potential has not yet been developed. Therefore, we should develop the interatomic potential describing the Mg–Mn binary system.
This study aims to develop the interatomic potential for the Mg–Mn binary system, as a prior study to investigate the dislocation behavior of the Mg–Mn alloy using the large-scale atomistic simulation, and finally reveal the effect of Mn on the RT formability of Mg. This paper begins by introducing the procedure to determine the potential parameters. Next, we confirm the reliability of the developed potential by computing fundamental material properties and comparing them with the data available from the literature.
Developing an interatomic potential means determining the model parameter values that compose the interatomic potential formalism for its associated system. The formalism used in this study is the second nearest-neighbor modified embedded-atom method (2NN MEAM) formalism,14) which has been widely used to develop interatomic potentials for Mg alloy systems.15–18) Under the 2NN MEAM formalism, a potential for a binary system comprises 13 binary model parameters, and 28 unary model parameters that compose the potentials for the two constituent unary systems. Each unary potential has 14 unary model parameters: cohesive energy (Ec), equilibrium nearest-neighbor distance (re), bulk modulus (B), a parameter (d) related to the pressure derivative of bulk modulus (∂B/∂P), decay lengths (β(0), β(1), β(2), β(3)), weight factors (t(1), t(2), t(3)), a parameter (A) for embedding function, and many-body screening parameters (Cmin, Cmax). The binary model parameters are Ec, re, B, d, the ratio between atomic electron density scaling factors (ρ0), four Cmin, and four Cmax. Detailed descriptions of the model parameters and the 2NN MEAM formalism are provided in Appendix A.
In this study, the 2NN MEAM potential for the Mg–Mn binary system was developed using the 2NN MEAM potentials for pure Mg18) and Mn (Table 1),19) which reproduce the fundamental material properties of their associated systems. The binary model parameters were determined to reproduce the fundamental material properties of the Mg–Mn system. Among the parameters, the Ec, re, B, and d parameter values constitute the universal equation of state for a selected reference structure. Thus, they can be determined from the corresponding experimental data for the reference structure if the structure is a stable compound in a phase diagram. The reference structure in the (2NN) MEAM formalism only can be a structure whose second nearest neighboring atoms are composed of the same element such as B1, B2, B3, and L12 structures. However, the Mg–Mn binary phase diagram does not contain any stable intermetallic compounds.6) Thus, we searched for information on metastable MgxMny compounds with these structures and found a first-principles calculation that reports the cohesive energy, lattice parameters, and bulk modulus of a hypothetical L12 Mg3Mn compound.20) Since the information was sufficient to determine the values of the Ec, re, and B parameters, we chose the L12 Mg3Mn compound as the reference structure of the Mg–Mn potential.
The model parameters for the Mg–Mn system were determined using the reference structures. The re parameter value was optimized to fit the lattice parameter of the Mg-rich hcp solid solutions. The Ec and B parameters were assigned values obtained from the first-principles calculation for the cohesive energy and bulk modulus of the hypothetical L12 Mg3Mn compound,20) respectively. The Cmin and Cmax parameter values were set from each assumption presented in Table 2. The value of the d parameter was set as the mean value of its constituent pure elements. The ρ0Mg:ρ0Mn ratio was given a default assumed value. The final potential parameter set determined for the Mg–Mn system is listed in Table 2.
In this section, to examine the reliability and transferability of the presently developed potential, the fundamental material properties of the Mg–Mn binary system are calculated and compared with data from the literature. The reliability was examined through properties related to the fitting quality. As mentioned before, there were no stable intermetallic compounds in the Mg–Mn phase diagram,6) so the fitting quality of the Mg–Mn system was examined by calculating the properties of stable solutions and hypothetical compounds. The transferability of the potential was examined to evaluate whether it could describe the dislocation behavior of the Mg–Mn alloy and could be used at finite temperatures.
Except for the enthalpy of mixing of liquid solutions and internal energy of compounds and solid solutions, the material properties were calculated at 0 K using in-house molecular statics and dynamics code, KISSMD,21) and a large-scale atomic/molecular massively parallel simulator (LAMMPS) package.22) A periodic boundary condition was set in all directions, except for the generalized stacking fault energy (GSFE) calculation. The radial cutoff distance was set to 6.0 Å which is a larger value than the second nearest-neighbor distances of Mg and Mn.
3.1 Fitting quality of the Mg–Mn potentialTo describe a system, an interatomic potential should reproduce the structural, elastic, and thermodynamic properties, which are the most fundamental material properties, of the stable phases in the system. In this respect, we first confirmed the reproducibility of the structural property of the presently developed Mg–Mn potential by calculating the lattice parameters of the Mg-rich hcp solid solutions owing to the lack of stable intermetallic compounds in the Mg–Mn system.6) For this calculation, we used eight random, disordered hcp solid solution samples containing 0 to 2.8 at% Mn, and calculated their lattice parameters at 0 K. Each sample contained 4000 atoms. The calculation results are presented in Fig. 1. The calculated lattice parameters decrease in value with the Mn content on both axes. The present calculations on the a-axis were consistent with the experimental data,23–25) whereas those on the c-axis were a little lower than the experimental data. Nevertheless, the decrement ratios of the calculated lattice parameters on both axes were in good agreement with the experimental data.
The reproducibility of the elastic property was confirmed by calculating the elastic moduli values of a hypothetical L12 Mg3Mn intermetallic compound, and the values are listed in Table 3, in comparison with a first-principles calculation.20) The calculated bulk modulus was consistent with the first-principles calculation.
The reproducibility of the thermodynamic property was confirmed by testing whether the present Mg–Mn potential could reproduce the Mg–Mn system that lacked any stable intermetallic compounds but contained a liquid miscibility gap at high temperatures.6) The absence of stable compounds was identified by calculating the enthalpies of formation of hcp and cbcc solid solutions and many hypothetical intermetallic compounds, and the values are presented in Fig. 2(a), in comparison with a first-principles calculation.20) The calculated enthalpy of formation of the hypothetical L12 Mg3Mn intermetallic compound was consistent with the first-principles calculation.20) Moreover, the calculated enthalpies of formation of all the hypothetical compounds were positive, which meant that they were unstable under the present Mg–Mn potential. Although the hypothetical compounds could not contain all the compounds in the whole materials, they contained the representative intermetallic compounds frequently formed in Mg alloys. Therefore, the present calculation results might reproduce the Mg–Mn system without stable compounds. Similar to the hypothetical compounds, the calculated enthalpies of formation of the hcp and cbcc solid solutions were all positive. This meant that the matrix in the Mg–Mn alloys was not composed of one kind of solid solution, i.e., the hcp solid solution or the cbcc solid solution, which is in agreement with the phase diagram of the Mg–Mn system whose matrix consists of the Mg-rich hcp solid solution and the Mn-rich cbcc solid solution.6) The enthalpies of mixing of liquid solutions were calculated to check the presence of the liquid miscibility gap using 13 random, disordered hcp solid solution samples containing 0–100 at% Mn. Each sample contained 4000 atoms. These solid samples were converted into liquid samples by raising their temperature to 2000 K. The liquid samples were cooled to 1600 K (as per temperatures in comparable literature).6,26) The calculation results are presented in Fig. 2(b), in comparison with CALPHAD calculations.6,26) While the general enthalpy of mixing shows parabolic curves, the calculated enthalpy of mixing appears to contain a straight line in the middle of the parabolic curve, which appears from 20 at% Mn to 90–95 at% Mn. This straight line is representative of a miscibility gab and also appears in the CALPHAD calculations.6,26) To check whether a phase separation occurs in the Mg–Mn liquid phase, we observed the atomic configuration of the Mg–50 at% Mn alloy simulated from the Parrinello/Rahman NPT ensemble at 1600 K for 250 ps. Figure 3 depicts that post-simulation, Mg and Mn atoms were divided into the Mg-rich liquid solution and the Mn-rich liquid solution. Although there is a discrepancy between the calculated enthalpy of mixing and the CALPHAD calculations (Fig. 2(b)), the present Mg–Mn potential could reproduce the phase separation phenomenon and reasonably predict the miscibility gap section.
(a) Enthalpies of formation of metastable hcp and cbcc solid solutions and hypothetical compounds at 0 K and (b) enthalpies of mixing of liquid solutions at finite temperatures of the Mg–Mn binary system according to the present 2NN MEAM potential, in comparison with a first-principles calculation20) and CALPHAD calculations.6,26)
Snapshots of (a) the initial and (b) final structures of the Mg–50 at% Mn alloy sample simulated from the Parrinello/Rahman NPT ensemble at 1600 K for 250 ps. Colors represent the element type of atoms (Green: Mg, Purple: Mn). The number of atoms in the sample is 4000.
In the previous section, we confirmed that the developed Mg–Mn potential can reproduce the structural, elastic, and thermodynamic properties of the stable phases in the Mg–Mn system. These properties are important for describing the alloy system but do not guarantee that the developed potential can reproduce the dislocation behavior of the Mg–Mn alloy. To confirm the reproducibility, we calculated the GSFE that is related to the dislocation behaviors of metallic alloys.27) The GSFE of pure Mg was also calculated for comparison. Among the stacking fault types in hcp metals, we investigated the I2-type stacking fault (…ABABCACA…), which is directly created by shear deformation. The calculation was conducted for representative slip planes and directions in hcp metals: basal $\{ 0001\} \langle 10\bar{1}0\rangle $, basal $\{ 0001\} \langle 11\bar{2}0\rangle $, prismatic $\{ 10\bar{1}0\} \langle 11\bar{2}0\rangle $, pyramidal I $\{ 10\bar{1}1\} \langle 11\bar{2}0\rangle $, pyramidal I $\{ 10\bar{1}1\} \langle 11\bar{2}3\rangle $, and pyramidal II $\{ 11\bar{2}2\} \langle 11\bar{2}3\rangle $ slip systems. The basal, prismatic, pyramidal I, and pyramidal II samples consist of 48 layers of the (0001) plane, 40 layers of the $(10\bar{1}0)$ plane, 36 layers of the $(10\bar{1}1)$ plane, and 40 layers of the $(11\bar{2}2)$ plane, respectively. The samples for the alloy system were created by replacing several Mg atoms on the stacking fault plane with Mn atoms. The solute concentration on the SF plane was 25.0 at%, which is equivalent to that from first-principles calculation28) with many comparable data. The detailed calculation procedure is described in our previous works.16,17) The calculated GSFE curves for the slip systems are shown in Fig. 4. The reliability of the calculated GSFE was verified by comparing their stable and unstable stacking fault energies (SFEs) with first-principles calculations28–32) (Table 4). The calculated stable SFE of the basal $\{ 0001\} \langle 10\bar{1}0\rangle $ slip system is larger in the order of pure Mg and Mg–Mn. The SFE order is consistent with those from first-principles calculations28,29) but is opposite to that of a first-principles calculation.30) The calculated stable SFE of the pyramidal II $\{ 11\bar{2}2\} \langle 11\bar{2}3\rangle $ slip system is larger in the order of pure Mg and Mg–Mn. Though there is no comparable data to confirm the consistency of this SFE order, these values are similar to those from a first-principles calculation.31) The calculated unstable SFEs of the basal $\{ 0001\} \langle 10\bar{1}0\rangle $ and $\{ 0001\} \langle 11\bar{2}0\rangle $ slip systems are both smaller in the order of pure Mg and Mg–Mn, which is opposite to the orders from the first-principles calculations.28,29) The calculated unstable SFE of the prismatic $\{ 10\bar{1}0\} \langle 11\bar{2}0\rangle $ slip system is larger in the order of pure Mg and Mg–Mn, which coincides with the SFE order from a first-principles calculation.28) The calculated unstable SFEs of the pyramidal I $\{ 10\bar{1}1\} \langle 11\bar{2}0\rangle $ and pyramidal I $\{ 10\bar{1}1\} \langle 11\bar{2}3\rangle $ slip systems are both larger in the order of pure Mg and Mg–Mn, and those of the pyramidal II $\{ 11\bar{2}2\} \langle 11\bar{2}3\rangle $ slip system are smaller in the order of pure Mg and Mg–Mn. The consistency of those SFE orders is difficult to confirm owing to the lack of comparable data, but those SEF values are similar to those from first-principle calculations.31,32)
Calculated GSFE (mJ/m2) curves of pure Mg and Mg–Mn, models on the (a) basal plane along the $\langle 10\bar{1}0\rangle $ direction, (b) basal plane along the $\langle 11\bar{2}0\rangle $ direction, (c) prismatic plane along the $\langle 11\bar{2}0\rangle $ direction, (d) pyramidal I plane along the $\langle 11\bar{2}0\rangle $ direction, (e) pyramidal I plane along the $\langle 11\bar{2}3\rangle $ direction, and (f) pyramidal II plane along the $\langle 11\bar{2}3\rangle $ direction, using the present 2NN MEAM potential.
In addition to the GSFE, we confirmed whether the presently developed Mg–Mn potential is reliable at finite temperatures for application in various future simulations. The reliability for the liquid phases was identified in the previous section (Fig. 2(b)), but that for the solid phases was not confirmed. To identify the reliability for the solid phases, we investigated whether the MgxMny hypothetical compounds and metastable solid solutions were unstable at finite temperatures. The stability of the compounds and solid solutions was identified using their internal energy after heating and rapid cooling to 0 K. Notably, the internal energy of the MgxMny intermetallic compounds and solid solutions should be less stable than the convex hull of pure Mg and pure Mn at the given temperatures and compositions. To confirm this, we heated the MgxMny hypothetical compound samples and the metastable hcp and cbcc solid solution samples to 1000 K and rapidly cooled them to 0 K from each heating temperature. The samples contained approximately 2000 atoms and the heating interval was 100 K. At each heating temperature, each sample was equilibrated for 30 ps. As shown in Fig. 5 and Fig. 6, the calculated internal energy of the whole MgxMny hypothetical compounds and metastable solid solutions was less negative than the convex hull, meaning that they were all unstable from 0 to 1000 K.
Change in the internal energy of stable (filled symbols) and metastable (open symbols) phases of (a) Mg3Mn, (b) Mg2Mn, (c) MgMn, (d) MgMn2, and (e) MgMn3 compounds after heating and rapid cooling to 0 K from each heating temperature.
Change in the internal energy of metastable hcp and cbcc solid solutions and hypothetical compounds in (a) Mg9Mn, (b) Mg8Mn2, (c) Mg7Mn3, (d) Mg6Mn4, (e) Mg5Mn5, (f) Mg4Mn6, (g) Mg3Mn7, (h) Mg2Mn8, and (i) MgMn9 compositions after heating and rapid cooling to 0 K from each temperature.
Taken together, the presently developed Mg–Mn interatomic potential can reproduce the structural, elastic, and thermodynamic properties of (meta)stable phases in the Mg–Mn system, and reasonably reproduce the GSFE of the Mg–Mn alloy. Furthermore, this potential can be used not only at 0 K but also at finite temperatures up to 1000 K. Therefore, the presently developed Mg–Mn interatomic potential can be utilized for studying the dislocation behavior of the Mg–Mn alloy, which will contribute to revealing the effect of Mn on the RT formability of Mg.
Based on the 2NN MEAM formalism, an interatomic potential for the Mg–Mn binary system was developed to infer the effect of Mn on the RT formability of Mg. The developed potential reasonably reproduced the structural, elastic, and thermodynamic properties of the compound and solution phases of the associated alloy system. This potential also described the GSFE of the basal, prismatic, pyramidal I, and pyramidal II slip systems of the Mg–Mn alloy, in reasonable agreement with first-principles calculations. Furthermore, it can be utilized from 0 K to 1000 K. The developed potential could be utilized to investigate the dislocation behavior of the Mg–Mn alloy and to elucidate the effect of Mn on the formability of Mg, which will further contribute to the enhancement of the RT formability and strength of Mg alloys.
This work was supported by the National Research Foundation of Korea (NRF) grants funded by the Korea government (MSIP) (No. 2021M3A7C2089767, 2021M3H4A6A01045764) and the Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Science, ICT & Future Planning (No. 2016R1A2B4006680).
In the MEAM, the total energy of a system is approximated as
\begin{equation} E = \sum_{i} \left[F_{i}(\bar{\rho}_{i}) + \frac{1}{2}\sum_{j(\neq i)} S_{ij}\phi_{ij}(R_{ij})\right] \end{equation} | (A.1) |
It has been shown in eq. (A.1) that the energy in the MEAM is composed of two terms, the embedding function term and the pair interaction term. The embedding function is given the following form,
\begin{equation} F(\bar{\rho}) = AE_{c}\frac{\bar{\rho}}{\bar{\rho}^{o}}\ln \frac{\bar{\rho}}{\bar{\rho}^{o}} \end{equation} | (A.2) |
\begin{equation} (\rho_{i}^{(0)})^{2} = \left[ \sum_{j\neq i} S_{ij}\rho_{j}^{a(0)}(R_{ij})\right]^{2} \end{equation} | (A.3a) |
\begin{equation} (\rho_{i}^{(1)})^{2} = \sum_{\alpha} \left[\sum_{j\neq i} \frac{R_{ij}^{\alpha}}{R_{ij}}S_{ij}\rho_{j}^{a(1)}(R_{ij})\right]^{2} \end{equation} | (A.3b) |
\begin{align} (\rho_{i}^{(2)})^{2} &= \sum_{\alpha,\beta} \left[ \sum_{j \neq i} \frac{R_{ij}^{\alpha} R_{ij}^{\beta}}{R_{ij}^{2}}S_{ij}\rho_{j}^{a(2)}(R_{ij}) \right]^{2}{}\\ &\quad - \frac{1}{3}\left[\sum_{j\neq i} S_{ij}\rho_{j}^{a(2)}(R_{ij})\right]^{2} \end{align} | (A.3c) |
\begin{align} (\rho_{i}^{(3)})^{2}& = \sum_{\alpha,\beta,\gamma} \left[\sum_{j\neq i} \frac{R_{ij}^{\alpha} R_{ij}^{\beta} R_{ij}^{\gamma}}{R_{ij}^{3}}S_{ij}\rho_{j}^{a(3)}(R_{ij})\right]^{2} \\ &\quad - \frac{3}{5}\sum_{\alpha} \left[\sum_{j\neq i} \frac{R_{ij}^{\alpha}}{R_{ij}}{S_{ij}}\rho_{j}^{a(3)}(R_{ij})\right]^{2} \end{align} | (A.3d) |
\begin{equation} \bar{\rho}_{i} = \rho_{i}^{(0)}G(\Gamma) \end{equation} | (A.4) |
\begin{equation} G(\Gamma) = \frac{2}{1 + {e^{-\Gamma}}} \end{equation} | (A.5) |
\begin{equation} \Gamma = \sum_{h=1}^{3} t_{i}^{(h)} \left[\frac{\rho_{i}^{(h)}}{\rho_{i}^{(0)}} \right]^{2} \end{equation} | (A.6) |
\begin{equation} \rho_{j}^{a(h)}(R) = \rho_{0}e^{-\beta^{(h)}(R/r_{e} - 1)} \end{equation} | (A.7) |
In the MEAM no specific functional expression is given directly to ϕ(R). Instead, the atomic energy (total energy per atom) is evaluated by some means as a function of nearest-neighbor distance. Then, the value of ϕ(R) is computed from known values of the total energy and the embedding energy, as a function of nearest-neighbor distance.
Let’s consider a reference structure once again. Here, every atom has the same environment and the same energy. If up to second nearest-neighbor interactions are considered as is done in the 2NN MEAM,14,36) the total energy per atom in a reference structure can be written as follows:
\begin{equation} E^{a}(R) = F(\bar{\rho}^{o}(R)) + \frac{Z_{1}}{2}\phi (R) + \frac{Z_{2}S}{2}\phi (aR) \end{equation} | (A.8) |
\begin{equation} E^{u}(R) = -E_{c}(1 + a^{*} + da^{*3})e^{-a^{*}} \end{equation} | (A.9) |
\begin{equation} a^{*} = \alpha (R/r_{e} - 1) \end{equation} | (A.10) |
\begin{equation} \alpha = \left(\frac{9B\Omega}{E_{c}}\right)^{1/2}. \end{equation} | (A.11) |
Basically, the pair potential between two atoms separated by a distance R, ϕ(R), can be obtained by equating eq. (A.8) and eq. (A.9). However, it is not trivial because eq. (A.8) contains two pair potential terms. In order to derive an expression for the pair interaction, ϕ(R), another pair potential, ψ(R), is introduced:
\begin{equation} E^{u}(R) = F(\bar{\rho}^{o}(R)) + \frac{Z_{1}}{2}\psi (R) \end{equation} | (A.12) |
\begin{equation} \psi (R) = \phi (R) + \frac{Z_{2}S}{Z_{1}}\phi (aR). \end{equation} | (A.13) |
\begin{equation} \psi (R) = \frac{2}{Z_{1}}[E^{u}(R) - F(\bar{\rho}^{o}(R))], \end{equation} | (A.14) |
\begin{equation} \phi (R) = \psi (R) + \sum_{n=1} (-1)^{n} \left(\frac{Z_{2}S}{Z_{1}} \right)^{n}\psi (a^{n}R). \end{equation} | (A.15) |
It should be noted here that the original first nearest neighbor MEAM is a special case (S = 0) of the present 2NN MEAM. In the original MEAM, the neglect of the second nearest-neighbor interactions is made by the use of a strong many-body screening function.35) In the same way, the consideration of the second nearest-neighbor interactions in the modified formalism is affected by adjusting the many-body screening function so that it becomes less severe. In the MEAM, the many-body screening function between atoms i and j, Sij, is defined as the product of the screening factors, Sikj, due to all other neighbor atoms k:
\begin{equation} S_{ij} = \prod_{k\neq i,j} {S_{ikj}} \end{equation} | (A.16) |
\begin{equation} x^{2} + \frac{1}{C}y^{2} = \left(\frac{1}{2}R_{ij} \right)^{2}. \end{equation} | (A.17) |
\begin{equation} C = \frac{2(X_{ik} + X_{kj}) - (X_{ik} - X_{kj})^{2} - 1}{1 - (X_{ik} - X_{kj})^{2}} \end{equation} | (A.18) |
\begin{equation} S_{ikj} = f_{c}\left[\frac{C - C_{\min}}{C_{\max} - C_{\min}}\right] \end{equation} | (A.19) |
\begin{equation} \begin{array}{ll} f_{c}(x) = 1 & x \geq 1\\ {[}1 - (1-x)^{4}]^{2} & 0 < x < 1\\ 0 & x \leq 0. \end{array} \end{equation} | (A.20) |
To describe a binary alloy system, in addition to the descriptions for individual elements, the pair interaction between different elements should be determined. For this, a similar technique used to determine the pair interaction for pure elements is applied. A perfectly ordered binary intermetallic compound, where one type of atom has only same type of atoms as second nearest-neighbors, is considered as a reference structure. The Bl (NaCl type) or B2 (CsCl type) ordered structure can be a good example. For such a reference structure, the total energy per atoms (for l/2 i atom + 1/2 j atom) is given as follows:
\begin{align} E_{ij}^{u}(R) &= \frac{1}{2}\bigg[ F_{i}(\bar{\rho}_{i}) + F_{j}(\bar{\rho}_{j}) + Z_{1}\phi_{ij}(R) \\ &\quad + \frac{Z_{2}}{2}( S_{i}{\phi_{ii}}(aR) + S_{j}\phi_{jj}(aR)) \bigg] \end{align} | (A.21) |
\begin{align} \phi_{ij}(R) &= \frac{1}{Z_{1}}\bigg[2E_{ij}^{u}(R) - F_{i}(\bar{\rho}_{i}) - F_{j}(\bar{\rho}_{j}) \\ &\quad - \frac{Z_{2}}{2}(S_{i}\phi_{ii}(aR) + S_{j}\phi_{jj}(aR))\bigg]. \end{align} | (A.22) |
The embedding functions Fi and Fj can be readily computed. The pair interactions ϕii and ϕjj between the same types of atoms can also be computed from descriptions of individual elements. To obtain $E_{ij}^{u}(R)$, the universal equation of state, eqs. (A.9)–(A.11), is considered once again for the binary reference structure. In this case, Ec, re and B correspond to the cohesive energy, equilibrium nearest-neighbor distance and bulk modulus of the binary reference structure.
In addition to the parameters for the universal equation of state, two more model parameter groups, Cmin and Cmax, must be determined to describe alloy systems. Each element has its own value of Cmin and Cmax. Cmin and Cmax determine the extent of screening of an atom (k) to the interaction between two neighboring atoms (i and j). For pure elements, the three atoms are all the same type (i-j-k = A-A-A or B-B-B). However, in the case of binary alloys, one of the interacting atoms and/or the screening atoms can be different types (there are four cases: i-k-j = A-B-A, B-A-B, A-A-B and A-B-B). Different Cmin and Cmax values may have to be given in each case. Another, the last model parameter necessary for binary alloy systems is the atomic electron density scaling factor ρo (see eq. (A.7)). For an equilibrium reference structure (R = re), the values of all atomic electron densities become ρo. This is an arbitrary value and does not have any effect on calculations for pure elements. This parameter is often omitted when describing the potential model for pure elements. However, for alloy systems, especially for systems where the constituent elements have different coordination numbers, the scaling factor (the ratio of the two values) has a great effect on calculations.