Next Article in Journal
Influence of Mold Design on Shrinkage Porosity of Ti-6Al-4V Alloy Ingots
Next Article in Special Issue
First-Principles Study on the Effect of H, C, and N at the Interface on Austenite/Ferrite Homojunction
Previous Article in Journal
Effect of Cu Additions on the Evolution of Eta-prime Precipitates in Aged AA 7075 Al–Zn–Mg–Cu Alloys
Previous Article in Special Issue
Phase Stability and Mechanical Properties Analysis of AlCoxCrFeNi HEAs Based on First Principles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Mechanical Properties, Structural Stability and Thermal Conductivities of Y, Sc Doped AuIn2 by First−Principles Calculations

Faculty of Materials Science and Engineering, Kunming University of Science and Technology, Kunming 650093, China
*
Author to whom correspondence should be addressed.
Metals 2022, 12(12), 2121; https://doi.org/10.3390/met12122121
Submission received: 10 October 2022 / Revised: 14 November 2022 / Accepted: 2 December 2022 / Published: 9 December 2022
(This article belongs to the Special Issue Application of First Principle Calculation in Metallic Materials)

Abstract

:
In this paper, based on density functional theory, the structural stability and mechanical properties of AuIn2 doped with RE (RE = Y, Sc) were investigated. The bulk modulus, shear modulus, Young’s modulus and Poisson’s ratio of the materials were calculated by Viogt−Reuss−Hill approximation. The calculation results show that Sc−SAu (trace of Au substituted by Sc in AuIn2), Y−SAu and Y−SIn have stable structure, and Y−SAu has obvious effect on the toughness indexes of AuIn2 alloy. Furthermore, based on Clarke and Cahill modes, the lattice thermal conductivity of the intermetallic compound was calculated and shows the same tendency with the Debye temperature and fast heat transfer rate in the direction of [110].

1. Introduction

AuM2 (M is Al, Ga, and In), the intermetallic compounds (IMCs) of stoichiometry AB2 with CaF2 type (fluorite) crystal structure [1,2], have been widely studied for their applications in wire bonding, chip packaging, aerospace, and medical equipment [3]. However, it is very brittle, fragile and difficult to process. Therefore, in order to obtain a new type of Au−based alloy with high strength and hardness to satisfy the requirements of structural and functional integration, a new type of Au alloy containing intermetallic compounds (IMCs) was developed, because IMCs can hinder the dislocation movement as the second phase in new Au−based alloy. The mechanical properties of IMCs are normally improved by adding trace elements. For example, B is added to Ni3Al [4] and Fe or Ga to NiAl [5], thus their room temperature plasticity is significantly improved. Meanwhile, the addition of trace elements can also reduce the melting point and increase the fluidity. Yu [6] calculated the mechanical propriety of AuAl2 by doping Y, Sc, Ga and In elements, and the experimental results show that Y, Sc doping has obvious toughening effect. Sinha [7] studied the phonon dynamics of AuAl2 and AuGa2 and found that there is a large deviation in their elastic constants although the bulk modulus of AuGa2 and AuAl2 alloys are very close. Rehmann S [8] demonstrated that AuIn2 doped with Y, Sc, Si displays a striking blue color, nuclear ferromagnetism and type I superconductivity. Godwal [9] found that AuIn2 starts to amorphize at 24 Gpa under high pressure and proved that the anomalous blue appearing at 2.7–2.8 Gpa is related to the content of Sc. Manjula [10] et al. studied the elastic and bond properties of zirconium and hafnium−doped Rh3V IMCs. The weaker covalent bond strength resulted in greater ductility, and Rh3Hf0.125V0.875 and Rh3Hf0.875V0.125 were more ductile. Furthermore, Manjula [11] also studied the relationship between valence−electron concentrations, formation energy and elastic mechanical properties of cubic Rh3AxTi1−x (A is V, Nb, Ta) and the results were further assessed by using charge density plots. Therefore, the doping of RE has crucial effects on the mechanical and thermal properties of the IMCs. Based on the first principle of Density Functional Theory (DFT), the structural stability, elasticity effect and thermal properties of AuIn2 doped with rare earth elements were studied in order to provide theoretical information for the further application of AuIn2 and related Au alloy.

2. Calculation Method

In this work, the CASTEP code [12] based on density functional theory [13] is used for the first−principles calculations [14]. According to the principle of selecting low energy, the CA−PZ scheme in the local density approximation (LDA) [15,16] is used to exchange related energy processing. The interaction between ion electronics and valence electrons were described by Ultrasoft pseudo−potentials (USPPs.) in which the valence electron configurations of Au are 4p45d106s1, In 4d105s25p1, Y 4d15s2 and Sc 3d14s2. To obtain the doping sites of RE (RE = Y, Sc), 2×2×2 supercells were built to ensure RE doping concentration of 1%, containing 32 Au atoms and 64 In atoms. The atomic positions and lattice parameters were optimized by the Broyden−Fletcher−Goldfarb−Shannon (BFGS) method. After the convergence test, the k point was set to 2×2×2 and cut−off energy was taken as 360 eV. During the structural optimization, the total energy convergence, the maximum stress and maximum force were set as 2 × 10−5 eV/atom, 0.1 GPa and 0.05 eV/Å, respectively.

3. Results and Discussion

3.1. Structural Properties and Formation Energy

As shown in Figure 1a, the crystal structure of AuIn2 is fluorite (CaF2) a cubic−type structure and belongs to the Fm3m space group. The AuIn2 crystal cell (lattice constant a = 6.502 Å) [17] possesses 4 Au atoms and 8 In atoms, where the Au atoms are located in (0,0,0)a and the In atoms in (1/4, 1/4, 1/4) a and (1/4, −1/4, −1/4) a. In atoms are located in the tetrahedral interstice location of face−centered cubic Au. Therefore, there are three possible RE doping sites (RE = Y, Sc) as shown in Figure 1c, namely octahedral interstice location (RE−I): (Au32REIn64), the substitution of Au atoms (RE−SAu: Au31REIn64) and the substitution of In atoms (RE−SIn: Au32In63RE). The optimized crystal lattice results of the doped and undoped systems are shown in Table 1. The deviations of the theoretically predicted AuIn2 lattice parameters from other theoretical values [18] and experimental data [19,20] are within 3%, pointing to the reasonable and feasible setting of the computational parameters in this paper. Compared with AuIn2, the crystal structure of the doped RE (RE = Y, Sc) systems show lattice distortions due to the difference of atomic radius between RE, Au and In. The lattice distortion caused by the replacement position is larger than that of the gap position, and the lattice distortion caused by Y doping is larger than that of Sc doping.
To understand the energy stability of rare earth element doping in neutral charge AuIn2, the formation energy (Ef) of three possible RE doping sites (RE = Y, Sc) was calculated by the following relation [21,22,23]:
E f ( D ) = E t o t a l ( D ) E t o t a l [ A u 32 I n 64 , b u l k ]   n i μ i
where Ef (D) denotes the formation energy of doped impurities in the supercell. Etotal (D) is the total energy of the supercell with doped impurities, while Etotal [Au32In64, bulk] is the total energy of Au32In64 and   μ i is the chemical potential of type i atoms. n i   states the number of type i atoms removed ( n i < 0) or added ( n i > 0) from the supercell during defect formation.
The formation energy of the Au atom substituted by RE (RE = Y, Sc) (RE−SAu: Au31REIn64) and the indium atom substituted by Y (Y−SIn: Au32YIn63) is negative, while the formation energy of RE (RE = Y, Sc) doped at the octahedral interstice location and indium atoms substituted by Sc (Sc−SIn: Au32ScIn63) system is positive. The high formation energy is accompanied by poor thermodynamic stability, so the structure of Y−SAu (Au31YIn64), as shown in Table 1, is the most stable. Stability was also certified by element electronegativity for the CaF2 type IMCs, because the electronegativity difference value between Au (2.54) and Y (1.22) is high compared to the difference of other elements; the bigger electronegativity difference value between two elements, the higher stability of the IMCs. The most stable structure is Au substituted by Y from the calculated results presented in Table 1.
In Table 2, the total bond population of the Au site and In site occupied by the RE atom obviously decrease, while the bond length of the doped systems increase. The higher the value, the stronger the bonding interaction and the greater the interatomic force. Therefore, the order of ductility of the IMCs is AuIn2 < Y−SIn < Sc−SAu < Y−SAu.

3.2. Elastic Constant of Single Crystal

According to Hooke’s law, the elastic constant was obtained from the stress–strain results [24]. In CASTEP code, each strain (ε) imposed on crystal would occur corresponding stress (σ), thus the step size of each strain is 4 and the maximum strain amplitude is 0.003 by the first−principal calculation. For different crystal structures, the number of independent elastic constants are determined by the structural symmetry. AuIn2 belongs to the cubic crystal system with three independent elastic constants C11, C12 and C44, as shown in Table 2.
To judge the structural stability, satisfaction of the corresponding innate mechanical stability criteria is also needed because different crystal structures have different stability criteria. For the cubic crystal system, the criteria of mechanical stability are as follows [25]:
C 11 > 0 , C 44 > 0 , C 11 > | C 12 | , ( C 11 + 2 C 12 ) > 0
The values of Cij in Table 2 are consistent with the above criterion of mechanical stability of cubic crystals, so it indicates that both AuIn2 and the RE doped are mechanically stable in the ground state. The linear compressive strengths of the cubic crystals along the X, Y, and Z axes are equal, due to C11 = C22 = C33 in the cubic crystals. In Table 2, the C11 of RE−SAu and Y−SIn are significantly lower than AuIn2, so the linear compressive strength of the doped system decreases along the X, Y and Z axes. In a cubic crystal, C12 represents the shear stress in the plane (110) along the direction of [ 110 ] . According to the values in Table 2, the C12 of RE−SAu and Y−SIn are greater than AuIn2, which indicated that the doped system has greater resistance to deformation in the (110) plane for the higher shear stress in the   [ 110 ] direction. Meanwhile, C44 represents the shear stress along the [001] direction on the (100) plane and the C44 of RE−SAu and Y−SIn is smaller than AuIn2, as shown in Table 2, which indicated that the doped system has less shear stress on the (100) plane in the [001] direction.

3.3. Elastic Modulus of Polycrystalline

The elastic modulus of polycrystals are normally obtained from the Voigt−Reuss−Hill approximate calculation model [26,27,28]. For cubic AuIn2 and its doping systems, the bulk and shear modulus of the Voigt, Reuss and Hill approximations were calculated by the following expressions [29]:
B V = B R = 1 3 ( C 11 + 2 C 12 )
G V = 1 5 ( C 11 C 12 + 3 C 44 )
G R = [ 4 5 ( C 11 C 12 ) 1 + 3 5 C 44 1 ] 1
B H = 1 2 ( B V + B R )
G H = 1 2 ( G V + G R )
Young’s modulus E and Poisson’s ratio ν are calculated as follows:
E = 9 B G 3 B + G
ν = 3 B 2 G 6 B + 2 G
The bulk modulus reflects the resistance of material to the external compression within the elastic range, namely the ability of material to resist deformation. The higher the bulk modulus, the higher the compressibility of the material. The values of the theoretical calculations of the elastic modulus, Poisson’s ratio and Cauchy pressure of AuIn2 and its doped systems are shown in Table 3. AuIn2 has the highest bulk modulus B 80.92 GPa, while that of the doped system is lower, which illustrates that the doped system has lower compressibility than AuIn2. The slightly different bulk modulus of the doping systems mean that RE−SAu (Y−SAu and Sc−SAu) have almost similar compression coefficients.
Shear modulus is the ratio of shear stress to shear strain within the range of the material elastic deformation. For cubic crystals, the shear modulus is related to the elastic constant C44 and a larger C44 corresponds to a larger shear modulus. The largest shear modulus of AuIn2 45.31 Gpa and the largest C44 55.4 GPa as shown in Table 3, indicating that the ability of AuIn2 to resist shear strain is stronger than that of the doped RE.
Young’s modulus is the ratio of stress to deformation within a physical elastic limit according to Hooke’s law. The higher the Young’s modulus, the higher the stiffness of a material, i.e., the material is less likely to deform. As Table 3 shows, the Young’s modulus E of AuIn2 is 117.23 Gpa, while the Young’s modulus of the doped system is significantly smaller than that of AuIn2. The order of Young’s modulus of the doping system is Y−SIn > Sc−SAu > Y−SAu.
Poisson’s ratio refers to the ratio of the absolute value of the transverse positive strain to the axial positive strain when the material is under unidirectional tension or compression. Poisson’s ratio was used to describe the strength of covalent bonds of the materials [30]. A smaller Poisson’s ratio is favorable to the formation of covalent bonds and a higher hardness. When Poisson’s ratio is in the range of −1.0 to 0.5, it means that the corresponding materials are stable [31]. The Poisson ratios of AuIn2, Sc−SAu, Y−SAu and Y−SIn are 0.264, 0.305, 0.320 and 0.290, respectively, and just within the range of −1.0 to 0.5. Therefore, AuIn2 and doping systems are stable in the linear elastic range.
The ratio of G/B, Cauchy pressure C12−C44 can also be sed to characterize the toughness and brittleness properties of cubic materials [32,33]. A material is brittle if it satisfies ν < 0.26, GH/BH > 0.57, C12−C44 < 0. Otherwise, the material is malleable. As Table 3 shows, GH/BH, ν and C12−C44 of AuIn2 doping systems meets the ductility condition, and the values of ν, C12−C44 for the Y−SAu system are the largest, while GH/BH is the smallest.
As Figure 2 shows, Cauchy pressure C12−C44 has the same trend as ν but the GH/BH shows an opposite trend. The Y−SAu system has the highest Cauchy pressure and ν but the lowest GH/BH which are especially different from AuIn2. This indicates that the earth elements doping (RE = Sc, Y) in AuIn2, especially the Y−SAu, improve the plasticity of the alloy.

3.4. Anisotropy of Elastic Modulus

The elastic anisotropy is one of the important mechanical properties of crystalline materials and is related to the generation of microcracks [32,34]. The value of elastic anisotropy AU = Ashear = 0 and A1 = 1 when the material is isotropic. Once the value of A deviates from 0, it shows that the crystal exhibits elastic anisotropy. The universal elastic anisotropy index (AU) [34], the anisotropy shear percentage (Ashear) [35] and the (100) plane shear factor (A1) [36] are expressed as follows:
A U = 5 G V G R + B V B R 6
A c o m p = B V B R B V + B R × 100 %    
A s h e a r = G V G R G V + G R × 100 %    
A 1 = 4 C 44 2 C 11 2 C 12
The calculated values of anisotropy indices AU, Ashear and A1 of AuIn2 and its doping system are shown in Table 4. The change trend of the anisotropy indices are shown in Figure 3.
The AU value 0.582 of Y−SAu and 0.202 of Y−SIn in Table 4 show that Y−SAu reached the maximum elastic anisotropy, and Y−SIn the minimum. The results can also be verified by Ashear which shows the same trend as AU of Y−SAu from Figure 3a, while Y−SIn has the smallest AU and Ashear values. The order of elastic anisotropy is Y−SAu > Sc−SAu > Y−SIn. This is consistent with the discussion of the bond length and formation energy in Table 1 and Table 2.
The values of shear anisotropy factor (A1) and the (|1−A1|) are used to characterize the shear anisotropy. The maximum value (0.934) of |1−A1| in Y−SAu indicates the highest shear anisotropy in (100), (010) and (001) planes, and the minimum (0.494) in Y−SIn the lowest in the same planes.
For cubic crystals, the anisotropy of bulk modulus B, shear modulus G and Young’s modulus E are calculated according to the following expressions (sij is elastic compliance constant) [37,38]:
1 B = ( s 11 + 2 s 12 ) ( l 1 2 + l 2 2 + l 3 2 )
1 G = s 44 + 4 ( s 11 s 12 s 44 2 ) ( l 1 2 l 2 2 + l 2 2 l 3 2 + l 1 2 l 3 2 )
1 E = s 11 2 ( s 11 s 12 s 44 2 ) ( l 1 2 l 2 2 + l 2 2 l 3 2 + l 1 2 l 3 2 )
The elastic anisotropy is related to the different atoms’ arrangement in the crystal structure. In order to observe the elastic anisotropy more intuitively, the elastic anisotropy is calculated according to the above expressions and directly characterized by three−dimensional (3D) surface structure, and the anisotropy increases with the increase of spherical deviation as shown in Figure 4, Figure 5 and Figure 6.
The bulk modulus B of AuIn2 and doping systems are isotropic because their 3D diagrams are almost spherical, as Figure 4 shows. This is consistent with the conclusion that the Acomp values of AuIn2 and the doping systems are zero. Au atoms replaced by RE (RE = Y, Sc) elements and In atoms by Y do not change the anisotropy of the bulk modulus, as Figure 4b–d shows.
In Figure 5, the deviation from sphere of the three−dimensional shear modulus G graph shows that the shear modulus G of AuIn2 and its doping system are anisotropic. The deviation is consistent with the maximum Ashear value of Y−SAu (5.49%). In addition, the larger deformation of Sc−SAu, but smaller than Y−SAu, is related to the order of anisotropy index Y−SAu > Sc−SAu > AuIn2. In contrast, the anisotropy index Ashear of Y−SIn is the smallest (1.99%) and the deviation of the three−dimensional graph from the sphere is small. Therefore, the order of shear modulus anisotropy is Y−SAu > Sc−SAu > AuIn2 > Y−SIn.
As Figure 6 shows, the section protruding along the vertex angle of the Y−SAu spatial distribution map displays a larger degree of deformation and higher degree of elasticity. In addition, the three−dimensional plots of Young’s modulus of Y−SAu and Sc−SAu are almost identical, which is related to their similar anisotropy index AU (Y−SAu: 0.582, Sc−SAu: 0.496), the same as AuIn2 and Y−SIn (AuIn2: 0.286, Y−SIn: 0.202). In contrast, the anisotropy index AU of Y−SIn is the smallest (0.202) and the almost round three−dimensional graph indicates that the elastic anisotropy of Y−SIn is the smallest. Therefore, the order of Young’s modulus anisotropy is Y−SAu > Sc−SAu > AuIn2 > Y−SIn, which is consistent with the anisotropic order from AU and Ashear.
To illustrate the elastic anisotropy of AuIn2 and its doped system in the (001) plane, we plotted the projections of AuIn2 and its doping system in the (001) plane with two major elastic modulus (bulk modulus and Young’s modulus), as shown in Figure 7. Meanwhile, the bulk modulus and Young’s modulus of AuIn2 doped with RE in each direction are listed in Table 5. For the cubic system, the atoms have the same arrangement in [100], [010] and [001] due to lattice symmetry, so they have the same elastic constants (B and E) along [100], [010] and [001]. In Figure 7a, the projections of bulk modulus of AuIn2 and the doping system are circular, thus the bulk modulus of AuIn2 and the doped systems are isotropic. In Figure 7b, the Young’s modulus of the model of Y−SAu displays the largest anisotropy. Moreover, in Table 5, the ratios of E[110]/E[100] of AuIn2, Sc−SAu, Y−SAu and Y−SIn are 1.06, 1.11, 1.18 and 1.02, respectively. The deviation order of E[110]/E[100] of the above models is Y−SAu > Sc−SAu > AuIn2 > Y−SIn.

3.5. Debye Temperature and Thermal Conductivity

The Debye temperature of the alloy were calculated according to the sound velocity ( υ m ). The formula is as follows [39,40,41]:
θ D = h k [ 3 n 4 π ( N A ρ M ) ] 1 3 υ m
υ m = [ 1 3 ( 2 v t 3 + 1 v l 3 ) ] 1 3 ,   v t = ( G ρ ) 1 2 ,   v l = ( 3 B + 4 G 3 ρ ) 1 2
where   v t   and v l are transverse and longitudinal sound velocities in the materials. h is Planck constant, approximately 6.626 × 10−34 J∙s; k is Boltzmann constant 1.38 × 10−23 J/K; n is the number of atoms; NA is Avogadro constant, 1.38 × 10−23 mol−1. ρ is the density and M is the molar mass of the material.
The longitudinal and transverse sound velocity, average sound velocity and Debye temperature θ D of AuIn2 including Y, Sc doped systems are listed in Table 6. It can be seen from Table 6 that the higher the density of the material, the higher the sound speed and Debye temperature. In contrast to AuIn2 with the highest density (6.196 g/cm3), Y−SAu shows the smallest Debye temperature (209 K), longitudinal speed of sound (4510 m/s), transverse speed of sound (2321 m/s), and mean speed of sound (1753 m/s).
The anharmonic of atomic interactions in crystalline solid materials is usually characterized by the acoustic Grüneisen constant γ a which is related to the heat transfer and expansion of the material and can be evaluated in terms of v l and   v t [41,42]:
γ a = 3 2 ( 3 v l 2 4 v t 2 v l 2 + 2 v t 2 )
The highest γ a (1.902) of Y−SAu and the lowest γ a (1.570) of AuIn2 in Table 6 indicate that the influence of external temperature and pressure on the lattice dynamics of the doping system are much greater than that of AuIn2.
The thermal conductivity will decrease to the limit as the temperature increases, thus the minimum thermal conductivity of materials can be used to evaluate the thermal conductivity ability in terms of crystalline solid materials [43]. The Clarke model [44,45] and Cahill model [46] were used to study the minimum lattice thermal conductivity k m i n of the IMCs. The Clarke model is described as follow:
k m i n = 0.87 k B M a 2 3 E 1 2 ρ 1 6
M a = [ M / ( m · N A ) ]
where M a   , ρ , M and m represent the average atomic mass, density, molar mass and total number of atoms of the unit cell, respectively.
Cahill’s model is described as follows:
k m i n = k B 2.48 n 2 3 ( v l + 2 v t )
where n is the number of atoms per unit volume.
In Table 7, the highest k m i n   of AuIn2 and the lowest value of Y−SAu are consistent with the highest Debye temperature of AuIn2 (243 K) and the lowest Debye temperature of Y−SAu (209 K) correspondingly. Thus, the minimum thermal conductivity coefficient k m i n shows the same tendency as the Debye temperature θ D because lower Debye temperatures correspond to lower lattice thermal conductivity [47].
In Clarke’s model, the k m i n of AuIn2 doped with RE is lower than that of AuIn2 because the atoms with heavy atomic mass are replaced by the lighter one. The order of atomic mass is Au > In > Sc > Y.
In Cahill’s model, the k m i n of AuIn2 is larger than the RE doped systems because of the phonon scattering effect caused by the atom replacement. Furthermore, the k m i n in Cahill’s model is higher than that in the Clarke model because of the neglect of phonons in the latter.
Based on the anisotropic of sound speed ( v l , v t 1 and v t 2 ), k m i n was used to characterize the anisotropy of the minimum thermal conductivity in the Cahill model:
k m i n = k B 2.48 n 2 3   ( v l + v t 1 + v t 2 )
As shown in Table 7, the k m i n [ 111 ] calculated by the Cahill model is extremely different in the directions of [100], [100] and [111]. The maximum value (0.901) of k min [ 110 ] and the minimum value (0.770) of k m i n [ 111 ] of AuIn2 indicate the faster heat transfer rates in the direction of [110], which are due to the closest atom alignment along the [110] direction in the cubic crystal, and are the same for the doping systems.

4. Conclusions

In this paper, we have explored the structure stability, elasticity and thermal conductivity of AuIn2 doped RE (RE = Sc, Y) by the first–principal calculations. According to the formation energies, together with the comparison to the element electronegativity difference value for the CaF2 type IMCs, the most stable doping of RE (RE = Sc, Y) in AuIn2 was Y−SAu (traces of Au atom were replaced by Y), but the octahedral interstice location doping was unstable.
The value of elastic constants of both AuIn2 and RE doped systems meet the criterion of mechanical stability and the doped systems have greater resistance to deformation along the [ 110 ] direction than along the [001] direction. Moreover, the higher Cauchy pressure and Poisson’s ratio but lower GH/BH of RE doping systems show that the RE elements in doping made the CaF2 type IMCs show ductile characteristics, of which the Y−SAu system is the most obvious. Furthermore, the increasing bond length of AuIn2 after RE doping is consist with the mechanical calculation results.
The anisotropic indexes of elastic anisotropy (AU, Acomp, Asheer, and A1), the 3D surface structure of the elastic modulus and the planar projection indicate that the anisotropic order of the elastic modulus is Y−SAu > Sc−SAu > AuIn2 > Y−SIn. Furthermore, the anisotropy of minimum thermal conductivity k min was calculated according to Clark and Cahill, and the results by Clarke models are in the same order as the corresponding Debye temperature, but opposite in terms of the elastic modulus. The maximum value of k min [ 110 ] and the minimum value of k min [ 111 ] are related to the different atomic mass from the Clarke model and to the different atom alignment, phonon scattering effect from the Cahill model.

Author Contributions

D.L.: Data curation; Formal analysis; Writing−original draft. J.L.: Investigation. Y.D.: Methodology. H.Q.: Validation. M.P.: Software. J.Y.: Writing—review & editing; Supervision; Resources. All authors have read and agreed to the published version of the manuscript.

Funding

National Natural Science Foundation of China (No. 51764030), the Major Special Project of Yunnan Province (202102AB080007); Natural Science Foundation of Yunnan Province (202001AS070010); Analysis and Testing Foundation of Kunming University of Science and Technology.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing is not applicable.

Conflicts of Interest

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

References

  1. Zintl, A.; Harder, A.; Haucke, Z.W. Legierungsphasen mit Fluoritstruktur. Phys. Chem. Abt. B 1937, 35, 354. [Google Scholar]
  2. Godwal, B.K.; Jayaraman, A.; Meenakshi, S.; Rao, R.S.; Sikka, S.K.; Vijayakumar, V. Electronic topological and structural transition in AuIn2 under pressure. Phys. Rev. B 1998, 57, 773–776. [Google Scholar] [CrossRef]
  3. Chi, X.P.; Xu, J.Y.; Zhong, S.P.; Chen, X.H. Research Progress of Microalloyed Gold−based Materials. Gold Sci. Technol. 2022, 30, 141–150. [Google Scholar] [CrossRef]
  4. Li, H.; Han, P.; Qi, Y.H. Toughening method and Mechanism of NiAlIntermetallics. J. Liaoning Inst. Technol. Nat. Sci. Ed. 2006, 26, 394–398. [Google Scholar]
  5. Xie, C.; Cheng, X.; Hu, G. Effects of iron added to NiAl intermetallic compound on the ductility and hot−rolling property. J. Mech. Eng. 1995, 31, 51–55. [Google Scholar]
  6. Yu, J.; Lu, J.K.; Cui, X.Y.; Ye, W.; Pu, R.; Zhou, X. Research Progress in Color and Technology of Au Based Materials for Jewelry. Mater. Rev. 2022, 12, 20100014-7. [Google Scholar] [CrossRef]
  7. Sinha, M.M.J. Phonon dynamics of binary intermetallic compounds AuX 2 (X = In, Al). J. Alloys Compd. 2010, 493, 577. [Google Scholar] [CrossRef]
  8. Herrmannsdörfer, T.; Rehmann, S.; Pobell, F. Interplay betweeen nuclear ferromagnetism and superconductivity in AuIn2. Czechoslov. J. Phys. 1996, 46, 2189–2190. [Google Scholar] [CrossRef]
  9. Godwal, B.K.; Speziale, S.; Clark, S.M.; Yan, J.; Jeanloz, R. Electronic phase transition and amorphization in AuIn2 at high pressure. Phys. Rev. B 2008, 78, 1884–1898. [Google Scholar] [CrossRef]
  10. Manjula, M.; Sundareswari, M.; Viswanathan, E. Elastic and thermodynamic properties of zirconium− and hafnium−doped Rh3V intermetallic compounds: Potential aerospace material. Bull. Mater. Sci. 2018, 41, 19. [Google Scholar] [CrossRef] [Green Version]
  11. Manjula, M.; Sundareswari, M.; Viswanathan, E. Enhancement of ductility in cubic Rh3A × Ti 1 − x(A = V, Nb, Ta)(x = 0, 0.125, 0.25, 0.75, 0.875, 1) aerospace materials–First principles DFT study. Mater. Chem. Phys. 2016, 181, 90–98. [Google Scholar] [CrossRef]
  12. Segall, M.D.; Philip; Lindan, J.D.; Probert, M.J.; Pickard, C.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]
  13. Kohn, W.; Sham, L.J. Self–consistent equations including exchange and correlation effects. Phys. Rev. 1965, 140, 1133–1138. [Google Scholar] [CrossRef]
  14. Kotsikau, D.; Ivanovskaya, M.; Orlik, D.; Falasconi, M. Gas–sensitive properties of thin and thick film sensors based on Fe2O3−SnO2 nanocomposites. Sens. Actuators B 2004, 101, 199–206. [Google Scholar] [CrossRef]
  15. Bond, G.C. The origins of particle size effects in heterogeneous catalysis. Surf. Sci. 1985, 156, 966–968. [Google Scholar] [CrossRef]
  16. Karazhanov, S.Z.; Ravindran, P.; Fjellvåg, H.; Svensson, B.G. Electronic structure and optical properties of ZnSiO3 and Zn2SiO4. J. Appl. Phys. 2010, 106, 123701–123701–7. [Google Scholar] [CrossRef] [Green Version]
  17. Zeitschrift fuer Physikalische Chemie, Abteilung B: Chemie der Elementarprozesse. Aufbau Der Mater. 1937, 35, 354–362. J. Less−Common Met. 1986, 116, 63–72.
  18. Wyckoff, R.W.G. New York Fluorite structure Crystal Structures, 2nd ed.; Interscience Publishers: New York, NY, USA, 1963; Volume 1, pp. 239–444. [Google Scholar]
  19. Warren, J.W.W.; Shaw, J.R.W.; Menth, A.; DiSalvo, F.J.; Storm, A.R.; Wernick, J.H. Nuclear Magnetic Resonance, Magnetic Susceptibility, and Lattice Constants of Solid and Liquid AuGa2, Au0.95Pd0.05Ga2, and AuIn2. Phys. Rev. B 1973, 2, 1247–1263. [Google Scholar] [CrossRef]
  20. Hiscocks, S.E.R.; Hume−Rothery, W. The Equilibrium Diagram of the System Gold−Indium. In Proceedings of the Royal Society of London. Series A, Mathematical and Physical Sciences, Online, 24 November 1964; Volume 282, pp. 318–330. [Google Scholar] [CrossRef]
  21. Zhang, S.; Northrup, J. Chemical potential dependence of defect formation energies in GaAs: Application to Ga self−diffusion. Phys. Rev. Lett. 1991, 67, 2339. [Google Scholar] [CrossRef]
  22. Laks, D.B.; Van de Walle, C.G.; Neumark, G.F.; Blöchl, P.E.; Pantelides, S.T. Native defects and self−compensation in ZnSe. Phys. Rev. B 1992, 45, 10965–10979. [Google Scholar] [CrossRef] [Green Version]
  23. Van de Walle, C.G.; Laks, D.B.; Neumark, G.F.; Pantelides, S.T. First–Principles Calculations of solubilities and doping limits: Li, Na, and N in ZnSe. Phys. Rev. B Condens. Matter 1993, 47, 9425–9434. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Wang, A.J.; Shang, S.L.; Du, Y.; Kong, Y.; Zhang, L.J.; Chen, L.; Zhao, D.D.; Liu, Z.K. Structural and elastic properties of cubic and hexagonal TiN and AlN from first−principles calculations. Comput. Mater. Sci. 2010, 48, 705–709. [Google Scholar] [CrossRef]
  25. Zhao, W.; Sun, Z.; Gong, S. Synergistic effect of co−alloying elements on site preferences and elastic properties of Ni3Al: A first−principles study. Intermetallics 2015, 65, 75–80. [Google Scholar] [CrossRef]
  26. Voigt, W. Handbook of Crystal Physics; Taubner: Leipzig, Germany, 1928. [Google Scholar]
  27. Reuss, A. Calculation of the flow limits of mixed crystals on the basis of the plasticity of monocrystals. Z Angew. Math. Mech. 1929, 9, 49–58. [Google Scholar] [CrossRef]
  28. Hill, R. The Elastic Behaviour of a Crystalline Aggregate. Proc. Phys. Soc. 1952, 65, 349–354. [Google Scholar] [CrossRef]
  29. Wu, Z.-J.; Zhao, E.-J.; Xiang, H.-P.; Hao, X.-F.; Liu, X.-J.; Meng, J. Crystal structures and elastic properties of superhard IrN2 and IrN3 from first principles. Phys. Rev. B 2007, 76, 059904. [Google Scholar] [CrossRef]
  30. Yao, B.X. First−Principles Calculation of the Stability and Properties of γ’−Pt3Al Phase by Alloying Element Doping; Kunming University of Science and Technology: Kunming, China, 2019. [Google Scholar]
  31. Bao, L.; Kong, Z.; Qu, D.; Duan, Y. Insight of structural stability, elastic anisotropies and thermal conductivities of Y, Sc doped Mg2Pb from first−principles calculations—ScienceDirect. Chem. Phys. Lett. 2020, 756, 137833. [Google Scholar] [CrossRef]
  32. Pettifor, D.G. Theoretical predictions of structure and related properties of intermetallics. Mater. Sci. Technol. 1992, 8, 345–349. [Google Scholar] [CrossRef]
  33. Pugh, S.F. XCII. Relations between the elastic moduli and the plastic properties of polycrystalline pure metals. Philos. Mag. 2009, 45, 823–843. [Google Scholar] [CrossRef]
  34. Ranganathan, S.I.; Ostoja−Starzewski, M. Universal Elastic Anisotropy Index. Phys. Rev. Lett. 2008, 101, 055504. [Google Scholar] [CrossRef] [Green Version]
  35. Vahldiek, F.W.; Mersol, S.A. Anisotropy in Single−Crystal Refractory Compounds; Springer: New York, NY, USA, 1968. [Google Scholar]
  36. Ravindran, P.; Fast, L.; Korzhavyi, P.; Johansson, B.; Wills, J.; Eriksson, O. Density functional theory for calculation of elastic properties of orthorhombic crystals: Application to TiSi2. J. Appl. Phys. 1998, 84, 4891–4904. [Google Scholar] [CrossRef]
  37. Nye, J.F. Physical Properties of Crystals; Wiley: New York, NY, USA, 1985. [Google Scholar]
  38. Hearmon, R. An Introduction to Applied Anisotropic Elasticity; Oxford University Press: Oxford, UK, 1961. [Google Scholar] [CrossRef]
  39. Music, D.; Houben, A.; Dronskowski, R.; Schneider, J.M. Ab initio study of ductility in M2AlC (M = Ti, V, Cr). Phys. Rev. B 2007, 75, 174102. [Google Scholar] [CrossRef]
  40. Chen, S.; Sun, Y.; Duan, Y.-H.; Huang, B.; Peng, M.-J. Phase stability, structural and elastic properties of C15−type Laves transition−metal compounds MCo2 from first−principles calculations. J. Alloy. Compd. 2015, 630, 202–208. [Google Scholar] [CrossRef]
  41. Okoye, C. Structural, elastic and electronic structure of LiCu2Si, LiCu2Ge and LiAg2Sn intermetallic compounds. Comput. Mater. Sci. 2014, 92, 141–148. [Google Scholar] [CrossRef]
  42. Arab, F.; Sahraoui, F.A.; Haddadi, K.; Bouhemadou, A.; Louail, L. Phase stability, mechanical and thermodynamic properties of orthorhombic and trigonal MgSiN2: An ab initio study. Phase Transit. 2016, 89, 480–513. [Google Scholar] [CrossRef]
  43. Shen, Y.; Clarke, D.; Fuierer, P. A Anisotropic thermal conductivity of the Aurivillusphase, bismuth titanate (Bi4Ti3O12): A natural nanostructured superlattice. Appl. Phys. Lett. 2008, 93, 102907. [Google Scholar] [CrossRef]
  44. Clarke, D.R. Materials selection guidelines for low thermal conductivity thermal barrier coatings. Surf. Coat. Technol. 2003, 163, 67–74. [Google Scholar] [CrossRef]
  45. Clarke, D.R.; Levi, C.G. Materials design for the next generation thermal barrier coatings. Annu. Rev. Mater. Res. 2003, 33, 383–417. [Google Scholar] [CrossRef]
  46. David, G.; Cahill, D.G.; Watson, S.K.; Pohl, R.O. Lower limit to the thermal conductivity of disordered crystals. Phys. Rev. B 1992, 46, 6131–6140. [Google Scholar] [CrossRef]
  47. Callaway, J. Model of lattice thermal conductivity at low temperatures. Phys. Rev. 1959, 113, 1046–1051. [Google Scholar] [CrossRef]
Figure 1. (a) AuIn2 unit cell, (b) 2 × 2 × 2 AuIn2 supercell, (c) RE doping sites in AuIn2 supercell. The golden balls are Au atoms and the pink balls are In atoms. The blue sphere represents the doped RE (RE = Y, Sc) atom and the positions 1, 2 and 3 represent the octahedral interstice location of RE doping, RE replacing In atom and RE replacing Au atom, respectively.
Figure 1. (a) AuIn2 unit cell, (b) 2 × 2 × 2 AuIn2 supercell, (c) RE doping sites in AuIn2 supercell. The golden balls are Au atoms and the pink balls are In atoms. The blue sphere represents the doped RE (RE = Y, Sc) atom and the positions 1, 2 and 3 represent the octahedral interstice location of RE doping, RE replacing In atom and RE replacing Au atom, respectively.
Metals 12 02121 g001
Figure 2. Variation of ν and C12−C44 (a), GH/BH (b) of AuIn2 and doped system.
Figure 2. Variation of ν and C12−C44 (a), GH/BH (b) of AuIn2 and doped system.
Metals 12 02121 g002
Figure 3. Variation of the elastic anisotropy indices AU, Ashear (a) and |1−A1| (b) of AuIn2 and doped systems.
Figure 3. Variation of the elastic anisotropy indices AU, Ashear (a) and |1−A1| (b) of AuIn2 and doped systems.
Metals 12 02121 g003
Figure 4. Three−dimensional surface structures of bulk modulus of AuIn2 and RE doped systems. For all the graphs, the units are GPa. (a) AuIn2 (b) Sc-SAu (c) Y-SAu (d) Y-SIn.
Figure 4. Three−dimensional surface structures of bulk modulus of AuIn2 and RE doped systems. For all the graphs, the units are GPa. (a) AuIn2 (b) Sc-SAu (c) Y-SAu (d) Y-SIn.
Metals 12 02121 g004
Figure 5. Three−dimensional surface structures of shear modulus of AuIn2 and RE doping systems; the units are GPa. (a) AuIn2 (b) Sc-SAu (c) Y-SAu (d) Y-SIn.
Figure 5. Three−dimensional surface structures of shear modulus of AuIn2 and RE doping systems; the units are GPa. (a) AuIn2 (b) Sc-SAu (c) Y-SAu (d) Y-SIn.
Metals 12 02121 g005aMetals 12 02121 g005b
Figure 6. Three−dimensional surface structures of Young’s modulus of AuIn2 and RE doped models. For all the graphs, the units are GPa. (a) AuIn2 (b) Sc-SAu (c) Y-SAu (d) Y-SIn.
Figure 6. Three−dimensional surface structures of Young’s modulus of AuIn2 and RE doped models. For all the graphs, the units are GPa. (a) AuIn2 (b) Sc-SAu (c) Y-SAu (d) Y-SIn.
Metals 12 02121 g006aMetals 12 02121 g006b
Figure 7. Projections of bulk and Young’s modulus of AuIn2 and its doped systems on (001) the crystal plane constant B (a), constant E (b).
Figure 7. Projections of bulk and Young’s modulus of AuIn2 and its doped systems on (001) the crystal plane constant B (a), constant E (b).
Metals 12 02121 g007
Table 1. Structural parameters and formation energy of AuIn2 and doped system.
Table 1. Structural parameters and formation energy of AuIn2 and doped system.
Phasea (Å)V3)Ef (eV)Refs
Au4In86.479271.972[18]
Au4In86.502274.879present
AuIn2 (4.2 K)6.483272.476[19]
AuIn2 (296 K)6.508275.640[19]
AuIn26.507275.513[20]
Au32In6412.9632178.29present
Sc−SAu12.9922192.95−1.61present
Sc−SIn12.9592176.28−0.85present
Sc−I13.1512274.454.99present
Y−SAu13.0142204.11−2.68present
Y−SIn12.9742183.84−1.15present
Y−I13.1672282.766.01present
Table 2. The calculated elastic constants Cij (in Gpa), the total number of population and bond length of AuIn2 and doped system.
Table 2. The calculated elastic constants Cij (in Gpa), the total number of population and bond length of AuIn2 and doped system.
PhaseC11C12C44Bond Population (Total)BondLength (Å)
AuIn2126.258.355.4138.24Au−In2.807
Sc−SAu111.763.045.6135.44Sc−In2.875
Y−Su111.766.543.7134.24Y−In2.998
Y−SIn121.159.845.8135.36Y−Au2.848
Table 3. The calculated bulk modulus B (in Gpa), shear modulus G (in Gpa), Young’s modulus E (in Gpa), Cauchy pressure C12−C44 (in Gpa), G/B ratio, and Poisson’s ratio ν of AuIn2 and doped system.
Table 3. The calculated bulk modulus B (in Gpa), shear modulus G (in Gpa), Young’s modulus E (in Gpa), Cauchy pressure C12−C44 (in Gpa), G/B ratio, and Poisson’s ratio ν of AuIn2 and doped system.
PhaseBGVGRGHEGH/BHC12−C44ν
AuIn280.9246.5744.0645.31117.230.562.90.26
Sc−SAu79.1837.0733.7235.3996.190.4517.40.31
Y−SAu80.8734.9531.3133.1391.660.4122.80.32
Y−SIn80.1039.9038.3439.11102.630.4914.00.29
Table 4. Calculated elastic anisotropic indexes (AU, Acomp, Ashear, and A1) of AuIn2 and doped systems.
Table 4. Calculated elastic anisotropic indexes (AU, Acomp, Ashear, and A1) of AuIn2 and doped systems.
PhaseAUAcompAshearA1
AuIn20.28602.771.632
Sc−SAu0.49604.731.873
Y−SAu0.58205.491.934
Y−SIn0.20201.991.494
Table 5. The bulk modulus and Young’s modulus of AuIn2 and doped systems in [100] and [110] directions (GPa) are calculated.
Table 5. The bulk modulus and Young’s modulus of AuIn2 and doped systems in [100] and [110] directions (GPa) are calculated.
PhaseBE
[100][110][100][110]
AuIn280.8780.8789.3294.89
Sc−SAu80.3280.3266.1673.4
Y−SAu110.12110.1262.1173.3
Y−SIn80.4580.4581.6086.56
Table 6. The density ρ (g/cm3), sound velocity (longitudinal v l (m/s), transverse v t (m/s) and mean v m (m/s), acoustic Grüneisen constant γ a and Debye temperature θ D (K) of AuIn2 and doping system.
Table 6. The density ρ (g/cm3), sound velocity (longitudinal v l (m/s), transverse v t (m/s) and mean v m (m/s), acoustic Grüneisen constant γ a and Debye temperature θ D (K) of AuIn2 and doping system.
PhaseAuIn2Sc−SAuY−SAuY−SIn
ρ 6.1966.1276.1476.185
v l 4776454145104624
v t 2704240323212515
v m 1941179417531849
γ a 1.5701.8071.9021.713
θ D 243216209227
Table 7. Calculated thermal conductivities k m i n ( W · m 1 · K 1 ) of AuIn2 and doping systems.
Table 7. Calculated thermal conductivities k m i n ( W · m 1 · K 1 ) of AuIn2 and doping systems.
PhaseClarke ModelCahill Model
M a   ( 10 22 )
k m i n  
k m i n [ 100 ]  
k m i n [ 001 ]
n (1028)
k m i n
k m i n [ 100 ]  
k m i n [ 110 ]  
k m i n [ 111 ]  
AuIn22.3620.4610.4030.4154.3590.7020.7230.9010.700
Sc−SAu2.3360.4200.3480.3674.3050.6390.6650.8300.642
Y−SAu2.3430.4090.3370.3664.2880.6240.6540.8190.632
Y−SIn2.3580.4320.3850.3904.3240.6620.6770.8600.661
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, D.; Lu, J.; Duan, Y.; Qi, H.; Peng, M.; Yu, J. The Mechanical Properties, Structural Stability and Thermal Conductivities of Y, Sc Doped AuIn2 by First−Principles Calculations. Metals 2022, 12, 2121. https://doi.org/10.3390/met12122121

AMA Style

Li D, Lu J, Duan Y, Qi H, Peng M, Yu J. The Mechanical Properties, Structural Stability and Thermal Conductivities of Y, Sc Doped AuIn2 by First−Principles Calculations. Metals. 2022; 12(12):2121. https://doi.org/10.3390/met12122121

Chicago/Turabian Style

Li, Deshuai, Jinkang Lu, Yonghua Duan, Huarong Qi, Mingjun Peng, and Jie Yu. 2022. "The Mechanical Properties, Structural Stability and Thermal Conductivities of Y, Sc Doped AuIn2 by First−Principles Calculations" Metals 12, no. 12: 2121. https://doi.org/10.3390/met12122121

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