Next Article in Journal
Deposition of CdSe Nanocrystals in Highly Porous SiO2 Matrices—In Situ Growth vs. Infiltration Methods
Previous Article in Journal
Severe Microbial Corrosion of L245 Transportation Pipeline Triggered by Wild Sulfate Reducing Bacteria in Shale Gas Produced Water
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Theoretical Analysis of Stacking Fault Energy, Elastic Properties, Electronic Properties, and Work Function of MnxCoCrFeNi High-Entropy Alloy

1
School of Intelligent Manufacturing Industry, Shanxi University of Electronic Science and Technology, Linfen 041000, China
2
School of Material Science and Engineering, North University of China, Taiyuan 030051, China
3
Department of Chemical and Materials Engineering, University of Alberta, Edmonton, AB T6G 1H9, Canada
*
Author to whom correspondence should be addressed.
Materials 2024, 17(17), 4378; https://doi.org/10.3390/ma17174378
Submission received: 23 November 2023 / Revised: 31 August 2024 / Accepted: 2 September 2024 / Published: 4 September 2024

Abstract

:
The effects of different Mn concentrations on the generalized stacking fault energies (GSFE) and elastic properties of MnxCoCrFeNi high-entropy alloys (HEAs) have been studied via first-principles, which are based on density functional theory. The relationship of different Mn concentrations with the chemical bond and surface activity of MnxCoCrFeNi HEAs are discussed from the perspectives of electronic structure and work function. The results show that the plastic deformation of MnxCoCrFeNi HEAs can be controlled via dislocation-mediated slip. But with the increase in Mn concentration, mechanical micro twinning can still be formed. The deformation resistance, shear resistance, and stiffness of MnxCoCrFeNi HEAs increase with the enhancement of Mn content. Accordingly, in the case of increased Mn concentration, the weakening of atomic bonds in MnxCoCrFeNi HEAs leads to the increase in alloy instability, which improves the possibility of dislocation.

1. Introduction

High-entropy alloys (HEAs), also known as multicomponent alloys, are alloys consisting of five or more equal or approximately equal amounts of metals. The concentration of each element in the alloy is between 5 and 35% [1,2]. It has a high entropy effect, extreme lattice distortion effect, cocktail effect, and sluggish diffusion effect [3,4,5,6]. In recent years, more and more people have paid attention to high-entropy alloys because of their unique compositions and excellent properties [7,8,9,10]. In particular, the high-entropy alloy Fe20Cr20Mn20Ni20Co20, with equal proportions, also known as the Cantor alloy, was first introduced in 2004 [11]. Therefore, this kind of CrMnFeCoNi alloy has long been a focus and hotspot of research.
So far, many researchers have extensively studied CoCrFeNi-based high-entropy alloys. Li et al. [12] designed a kind of metastable high-entropy dual-phase alloy, Fe80-xMnxCo10Cr10. They concluded that the single face-centered cubic (FCC) phase structure could be satisfied when the Mn content is 40 and 45% (Fe35Mn45Co10Cr10 and Fe40Mn40Co10Cr10, respectively). The Mn content plays an important role in phase composition, regulating phase stability and improving phase transition mechanism. Sun et al. [13] compared the lattice stability of AlxCrMnFeCoNi and AlxCrFeCoNi high-entropy alloys. It was found that Mn decreases the stable field of FCC phase and widens the width of the two-phase region. Moreover, via the exact muffin-tin orbitals (EMTO) method, Zhang et al. [14] calculated the elastic properties of body-centered cubic (BCC) and face-centered cubic (FCC) AlxCrMnFeCoNi (0 ≤ x ≤ 5) HEAs, demonstrating that there is a complex dependence between the elastic parameters and the composition, and the elastic anisotropy of both phases is extremely high. Shi Y.Z. et al. [15] studied the homogenization effect of 1250 °C heat treatment on AlxCoCrFeNi HEAs. After heat treatment, the homogenization effect of microstructure leads to the decrease in work function and the improvement of corrosion resistance. Zhang et al. [16] prepared CoCrFeNi-Nbx (x = 0, 1, 3, 5, 7, 9 wt%) high-entropy alloys through high-energy ball milling and discharge plasma sintering. The effects of niobium on the microstructure and properties of cobalt–nickel alloys were studied systematically. Nb atoms cause lattice distortion of the alloy, and the microstructure of CoCrFeNi HEAs changes from a single-phase structure of FCC to a bi-phase structure of FCC and Laves, which increases the tensile strength, yield strength, and hardness of the HEA.
Moreover, Kivy M.B. et al. [17] investigated the effects of Cu, Mn, Al, Ti, and Mo on the generalized stacking fault energies (GSFE), Rice-criterion ductilities, and twin ability of CoCrFeNi-based HEAs with FCC structure via density functional theory. The results presented that the addition of Ti and Mo increases the tendency of dislocation slip and deformation twinning. Furthermore, the addition of Mn, Cu, or Al with high content promoted dislocation slip and martensitic phase transition, while a low amount of Al led to dislocation slip. Achmad et al. [18] have calculated pure cobalt and Co-9 at.% X solid-solution alloys (X = Cr, W, Mo, Ni, Mn, Al, Fe) using first-principles density-functional theory. They found that the alloying of Cr, W, and Mo increases the interlayer distance distortion of the stacking fault plane and reduces the GSFE value of pure Co due to the increase in charge accumulation. Also, the stacking fault energy (SFE) values of several typical FCC high-entropy alloys (HEAs) were measured by Liu et al. [19] using experimental means. The experimental results indicated that the lower SFE is helpful for the formation of deformation twins under the loading condition, and the smaller the thickness is, the better the mechanical properties are at low temperature.
It is obvious from the above research results that the Mn element plays an important role in the CoCrFeNi-based high-entropy alloy. Although a great deal of research has been conducted on the effects of alloying elements, especially Mn, on high-entropy alloys, there is still a lack of theoretical research on the GSFE and work function of CoCrFeNi-based high-entropy alloys with different Mn. In general, the GSFE has been widely studied in characterization of the mechanical properties and brittle-to-ductile transition of alloys. The GSFE is considered to be a measure of energy penalty between two adjacent planes during shear deformation along a specific slip direction on a given slip plane. It represents the nature of the slip and involves stable and unstable stacking and twin fault energy. In particular, the intrinsic stacking fault energy (ISF) of the alloy can be calculated via transmission electron microscopy (TEM) [20,21] and X-ray powder diffraction (XRD) [22]. However, the unstable stacking fault energy (USF) cannot be measured by experiment, and the first principle calculation is an effective method by which to measure the USFE [23]. On the other hand, virtual crystal approximation (VCA) is considered a method by which to study the properties of solid solutions. The method uses “virtual” atoms inserted between atomic behaviors in the original compound to study crystals in primitive periodicity. This method makes the calculation simpler and the cost lower. Of course, previous work [24,25] has proved that VCA has good accuracy in some HEAs.
Therefore, based on the first-principle calculation, the theoretical models of MnxCoCrFeNi HEAs are established via the VCA method in the present work. In order to study the effect of different Mn concentrations on slip and twin in MnxCoCrFeNi HEAs, 13 close-packed (111) atomic layers are used to calculate GSFE, as well as USF, ISF, unstable twin-fault energy (UTF), and two-layers twinning stacking fault energy (TSF) parameters. More importantly, the elastic properties and electron density difference were systematically studied with the aim of further uncovering the nature of the strength and ductility of MnxCoCrFeNi HEAs. Finally, the work functions of (111), (110), and (100) planes in MnxCoCrFeNi HEAs are estimated, which reveals the analysis results in more details. Therefore, theoretical analysis was conducted on HEAs containing different proportions of Mn elements in this study. The influence of Mn element on the GSFE and other properties of CoCrFeNi-based HEAs can be obtained, which can provide useful guidance for the design of high-performance CoCrFeNi-based HEAs.

2. Methods and Details

2.1. First-Principles Calculations

All the first-principles calculations are performed using the Cambridge sequential total energy package (CASTEP) (MS 7.0) [26,27] based on density functional theory (DFT) [28,29]. The exchange correlation functional uses the Pardew–Burke–Ernzerhof (PBE) generalized gradient approximation (GGA) [30,31]. Clearly, the premise for simulation is geometry optimization (minimum energy of atoms at different volumes). Therefore, according to the termination of structural relaxation, the convergence parameters are selected. The convergence parameters were set as follows: total energy tolerance is 10−5 eV/atom; force tolerance is 0.03 eV/Å; maximum stress is 0.05 GPa; and maximum displacement is 0.001 Å. After the convergence test, the plane wave energy cutoff is 600 eV and the Monkhorst–Pack scheme [32,33] k-points set is 10 × 10 × 10 in the Brillouin zone. According to the selected setting, the error reaches less than 1% through the convergence of the total energy of the calculated model.

2.2. VCA Models

The construction of VCA unit cells is achieved by replacing real alloy atoms in the cell structure with “virtual” atoms, which are obtained from the weighted average technique of different alloy elements. The atomic percentages of each constituent atom within the cell of the MnxCoCrFeNi high-entropy alloy are presented in Table 1.
Accurately determining the lattice constant of the unit cell is an essential prerequisite for predicting the properties of compounds in theoretical calculations. In this paper, the comparison between the optimized lattice constant and the lattice constant obtained from other calculation methods and experimental results is detailed in Table 2. In Table 2, SQS and CPA represent the estimation of lattice constants of high-entropy alloys using a special quasi random structure (SQS) and coherent-potential approximation (CPA), respectively. As shown in Table 2, the calculated lattice constant is consistent with other calculation methods and experimental results.

2.3. Stacking Fault Model and GSFE

The structure diagram of the FCC structural unit along [111], [ 1 ¯ 10], and [11 2 ¯ ] directions is shown in Figure 1a. Based on the fact that dislocations in the FCC structure mainly occur on the (111) surface of dense packing, a supercell model with 13 closed-stacking (111) is constructed, in which the stacking sequence from bottom to top is “ABCABCABCABCA”. In Figure 1b, the GSFE value was obtained vis shear slip along [11 2 ¯ ] in the plane (111). In the first process, the upper 7–13 layers of atoms are sheared along the [11 2 ¯ ] direction. With the step size of 0.1bp, the unstable stacking fault (USF) is generated at the shear displacement of 0.0 to 0.5 bp, in which the Burgers vector b p = a 0 / 6 . When the shear displacement increases to 1.0 bp, the structure becomes the ISF of ABCABCBCABCAB sequence. The second process is to move the upper 8–13 layers of atoms at the same distance along the [11 2 ¯ ] direction. When the shear displacement value increases to 1.0 bp, the structural sequence becomes ABCABCBABCABC, which is the TSF. The GSFE is calculated as follows [18]:
γ G S F E = 1 A E u E 0 ,
where Eu is the total energy of the supercell after shear displacement, E0 is the total energy of the supercell without defect, and A is the area of the fracture plane. To avoid periodic atomic interactions, a 15 Å vacuum layer is added to the structure. After adequate convergence testing for the GSFE calculation, and considering the calculation efficiency, the plane wave energy cutoff and the k-points are set as 400 eV and 10 × 10 × 1, respectively.

2.4. Surface Energies

In order to better understand the surface properties of MnxCoCrFeNi HEAs, the surface energies (γs) of (100), (110), and (111) planes were calculated. Figure 2 illustrates the (100), (110), and (111) surface models of MnxCoCrFeNi HEAs, which contain 10 atomic layers. A vacuum layer with a thickness of 5 Å is set on the bottom and top of the surface model to eliminate the influence of the interaction between periodic structures, as shown in Figure 2a,c,e. With the aim of avoiding the interaction between atoms in the slab, an 8 Å vacuum layer is set between the atomic layers of the slab model (Figure 2b,d,f). The surface energy can be calculated by the following formula [25,37]:
γ s = 1 2 A E s l a b E B u l k .
here, E s l a b is the total energy of the surface model. E B u l k is the total energy of the perfect supercell, and A is the surface area.

3. Results and Discussion

3.1. GSFE and Surface Energies

Figure 3 and Table 1 exhibit the GSFE curves and related values of MnxCoCrFeNi HEAs with different Mn concentrations, which were calculated based on first-principles. According to the Figure 3, with the increase in shear displacement along the direction of [11 2 ¯ ], the structure exhibits an energy barrier at the first highest energy point on the GSFE curve, which is called the unstable stacking fault energy (γusf). This energy barrier is considered as the minimum energy or critical stress required for local dislocation nucleation. As the structure is further sheared, some dislocations begin to spread, and a stacking fault defect is created. The first minimum energy point on the GSFE curve is formed, and this energy is called intrinsic or stable stacking fault energy (γisf). The second maximum energy point on the GSFE curve is called the unstable twin fault energy (γutf). This energy is considered to be the minimum energy barrier to produce an extrinsic or twinning stacking fault. The second minimum energy point is defined as the extrinsic stacking fault energy or two-layer twinning stacking fault energy (2γtsf). The GSFE calculation based on the first-principle separates the formation energy from any other related effects on total energy, so the calculated value may be more accurate than the experimental measurement value [18]. Van Swygenhoven et al. [38] found that only using SFE parameters as a criterion for judging the twinning and partial dislocation deformation mechanism of materials is not sufficient.
As can be seen from the Figure 3, with the increase in Mn concentration, the GSFE curves of MnxCoCrFeNi HEAs gradually decreased. The USF energy decreases from 808.84 mJ/m2 to 276.17 mJ/m2 with increasing Mn content from 0.0 to 1.0. The results demonstrate that the alloy with higher Mn content easily forms dislocation due to its low nuclear dislocations resistance. The two-layer twinning SFE of HEAs was lower than that of the stable SFE. And the decrease degree of the two-layer twinning SFE was greater than that of stable SFE. When the Mn content was 0, the two-layer twinning SFE of HEAs (2γtsf = 574.71 mJ/m2) was lower than the stable SFE (γisf = 578.93 mJ/m2). When the Mn content was 1, the value of the two-layer twinning SFE is 60.36 mJ/m2, and the value of the stable SFE is 110.04 mJ/m2.
The ratio of ISF energy to USF energy (γisfusf) is often used to indicate the tendency of complete dislocation dissociation into partial dislocation. The lower the ratio, the greater the tendency of full dislocation dissociation. When the leading partial dislocations nucleate by overcoming the energy barrier γusf, the tailing part needs to exceed this energy for nucleation. The critical stress for the nucleation of the trailing part is a function of (γusf − γisf). The increased value of γusf − γisf is beneficial to the partial dislocation, and the stacking fault more likely to form. The difference between γusf and γisf in MnxCoCrFeNi HEAs was large, implying that the deformation mechanism was only extended partial dislocation. Moreover, with the increase in Mn content, the difference value decreased gradually from the original 229.91 to 166.13 mJ/m2. The result displays that the tendency of stacking faults decreases with the increase in Mn content. As indicated in Figure 4a, the ratio of γisf and γusf in MnxCoCrFeNi HEAs was larger, demonstrating that the energy barrier needed to form the trailing part was smaller, and it was difficult to produce stacking faults. In addition, Tadmor et al. [39] gave a standard by which the main deformation mechanism can be judged when mechanical twinning becomes an ideal crack tip. By this method, the tendency of partial dislocations to form total dislocations can be determined, resulting in dislocation-mediated slip or mechanical twins. Its expression is as follows.
δ u s f u t f = γ u t f γ u s f ,
It is clear from Table 3 that the δ u s f u t f values of MnxCoCrFeNi HEAs are positive, suggesting that the energy barrier for unstable twin formation (γutf) is greater than the energy barrier for partial dislocation propagation (γusf). It can be considered that the deformation mechanism of MnxCoCrFeNi HEAs is not conducive to the transformation from dislocation to twins. This conclusion suggests that plastic deformation can be dominated by dislocation-mediated slip. However, with the increase in Mn concentration, its value decreases. It is hinted that mechanical twin can still be formed with the increase in Mn concentration. The existence of twin propagation starts from local dislocation or twin crystals, which is related to the γutfusf ratio. If the ratio is low, twin deformation is more likely to occur. It is obvious in Table 1 that the ratio of γutfusf has little correlation with the change in Mn concentration. For further study, Tadmor et al.’s [39] criterion was used to study the tendency of FCC metal forming mechanical twins. The formula is as follows:
τ a = 1.136 0.151 γ i s f γ u s f γ u s f γ u t f ,
Among them, 1.136 and 0.151 are the general coefficients of the FCC lattice. A higher τ a value can achieve higher twinning tendency. According to Figure 4b, the τ a values increase with the increase in the concentration of Mn in MnxCoCrFeNi HEAs, but the change in τ a was little. When the Mn concentration increases from 0.0 to 1.0, the τ a value increases from 0.956 to 0.969, and the difference is only 0.013. It follows that the increase in Mn concentration can lead to twin formation, but the effect is not significant.
An alternative parameter τ, given by Asaro et al. [40], can reflect the competition between dislocation propagation and grain boundary source mechanical twins, and the expression is expressed as follows:
τ = 1 + 2 β γ u s f γ u t f ,
where β = 1 γ i s f / γ u s f . When τ > 1, the twin phenomenon is considered to be the dominant mechanism, which is more conducive to complete dislocation. It can be observed from Figure 4c that all the τ values of HEAs are greater than 1, and the value gradually increases from 1.164 to 1.337 with the increase in Mn concentration. This shows that the increase in Mn concentration is more favorable for twin generation.
Kibey et al. [41] proposed a continuous, multi-scale method by which to predict twin stresses. The method uses the dislocation-based twin nucleation model to calculate the critical twin stress of GSFE in FCC metal, and the expression is as follows:
τ c r i t = 5 18 b t w i n γ u t f + 2 γ t s f + γ i s f 2 2 9 b t w i n γ u s f + γ i s f .
Twin phenomena in FCC structures are caused by shear along the direction of [11 2 ¯ ] on the (111) surface, so btwin is defined as a / 6 . According to Equation (6), τ c r i t is determined by the four typical GSFE values (γisf, γusf, γutf, and 2γtsf), instead of only considering the correlation of γisf. This method has been used to predict twin stresses of some FCC metals. When the τ c r i t value decreases, the tendency to produce twins increases. Figure 4d demonstrates that the τ c r i t value decreases gradually with the increase in Mn content, from 77.83 to 23.20. It is obvious that alloys with higher Mn content require less energy for critical twinning stress than alloys with lower Mn content, and twins are more likely to be produced.
The influence of different Mn contents on the ductility of MnxCoCrFeNi HEAs can be analyzed using the Rice-criterion [42]. This analysis explains the competitive relationship between crack tip dislocation formation and crack cleavage. Its expression can be expressed as follows:
D = 0.3 γ s γ u s f ,
where D is the ductility parameter and γs is the surface energy along the direction of [111]. When D > 1, the dissociation energy of the crack is greater than that of the dislocation nuclear energy. As a result, the alloy will exhibit ductile behavior. However, in the event of D < 0.3, the failure was caused by crack cleavage rather than dislocation slip. The calculated results listed in Figure 4e and Table 3 indicate that the D values of all alloys are greater than 1 and increase continuously with the increase in Mn concentration. The results show the MnxCoCrFeNi HEAs’ ductility, and there was dislocation formation of the crack tip. Kivy M.B. et al. [17] studied the Rice-criterion ductilities for the alloy (D > 1), indicating that the display ductility of the alloy is consistent with the calculation results in this paper.
The calculated surface energies of MnxCoCrFeNi HEAs (111), (110), and (100) surfaces are summarized in Figure 4f. It is obvious that in HEAs with different Mn concentrations, the close-packed (111) plane has the lowest surface energy. With the increase in Mn content, the surface energy of the three planes tends to decrease. When the x value changes from 0.1 to 1.0, in MnxCoCrFeNi HEAs, the surface energy differences of (111), (110), and (100) is 1.447 J/m2, 1.102 J/m2, and 1.071 J/m2, respectively. Therefore, the surface energy of surface (111) is most susceptible to the increase in Mn concentration.

3.2. Elastic Properties

With the goal of studying the elastic properties of MnxCoCrFeNi HEAs, the elastic constants of the alloys were also calculated. Furthermore, the elastic modulus, Poisson’s ratio, and anisotropy of the alloys are analyzed. MnxCoCrFeNi HEAs are a cubic crystal system, which has three independent elastic constants: C11, C12, and C44. The change in elasticity constant of MnxCoCrFeNi HEAs with Mn concentration is expressed in Figure 5a. As displayed in the Figure 5a, the elastic constant of the alloy conforms to the mechanical criterion of the cubic crystal system [43]: C11 > 0; C44 > 0; C11−C12 > 0; C11 + 2C12 > 0. This is evidence that these alloys are mechanically stable. The three elastic constants gradually increase with the increase in Mn content. The change in C11 was more significant. As the value of x increases from 0 to 1, the value of C11 increases from 237.37 GPa to 357.48 GPa. Compared with C11, the variation trend of C12 and C44 is relatively gentle.
The bulk modulus (B), shear modulus (G), Young’s modulus (E), Vickers hardness (Hv), and Poisson’s ratio (σ) of the alloy can be calculated according to the following formulae [44,45]:
B H = 1 2 B V + B R ,
G H = 1 2 G V + G R ,
E = 9 B H G H ( 3 B H + G H ) ,
H v = 2 G 3 / B 2 0.585 3 ,
σ = ( 3 B H 2 G H ) 2 3 B H + G H ,
where subscripts V and R are the approximate values obtained by Voigt and Reuss, respectively. Subscript H represents the average values of these two approximations. The bulk modulus reflects the ability of the alloy to resist bulk deformation. Shear modulus can be used to measure the shear deformation resistance of the alloy. Young’s modulus can reflect the stiffness of the alloy. As can be seen from Figure 5b, with the increase in Mn concentration, the bulk modulus of HEAs increases gradually. The bulk modulus of the HEAs gradually increases from 188.23 GPa to 267.280 GPa. The change trend of MnxCoCrFeNi HEAs’ shear modulus and Young’s modulus was consistent with the bulk modulus. When x is 0.7, the shear modulus and Young’s moduli of HEAs reach 109.951 GPa and 288.592 GPa, respectively. As the concentration of Mn continues to increase, the change trend tended to be gentle. The results demonstrate that along with the raise in Mn content, the volume deformation resistance, shear deformation resistance, and alloy stiffness of HEAs increases continuously. The Vickers hardness can be macroscopic reaction alloy hardness. It can be seen from Equation (11) that Vickers hardness is affected by bulk modulus and shear modulus. It can be observed from the Figure 5b that the Vickers hardness of the alloy increases with the increase in Mn concentration. When the value of x is 0.7, the Vickers hardness of the alloy achieves 8.615 GPa. After that, with the increase in Mn concentration, the Vickers hardness of the HEAs is almost unchanged. It is obvious that when Mn content is greater than 0.7, the Vickers hardness is no longer affected by Mn content.
The critical value of the Poisson’s ratio, which is 0.26, is often used to judge the ductility and brittleness of alloys. When the Poisson’s ratio is greater than 0.26, the compound is ductility. On the contrary, when the Poisson’s ratio is less than 0.26, the compound is brittle. It can be visualized in Figure 5c that the Poisson’s ratios of MnxCoCrFeNi HEAs were all greater than 0.26, hinting that these alloys are ductile. B/G is another criterion for judging the brittleness and ductility of alloys. According to Pugh criterion [46], if the B/G value of an alloy is greater than 1.75, the material has ductility. In it is not, the material is brittle. It is presented in the Figure 5c that the values of B/G are all greater than 1.75, and their variation trend is similar to that of Poisson’s ratio. This is evidence that the alloys are ductile, which is consistent with the Poisson’s ratio analysis.
The anisotropy of crystal elasticity plays an important role in the study of macroscopic mechanical properties. Therefore, it is of great significance to study the influence of anisotropy of MnxCoCrFeNi HEAs on the mechanical behavior of the alloy. The anisotropy index (AB and AG) and the general anisotropy index (AU) can be calculated using the following formulae [47]:
A B = B V B R B V + B R × 100 % ,
A G = G V G R G V + G R × 100 % ,
A U = 5 G V G R + B V B R 6 0 .
When the values of AB, AG, and AU are zero, this indicates that the crystal appears isotropic. On the contrary, when the anisotropy index is not equal to zero, the crystal shows anisotropy. The greater the deviation from 0, the greater the anisotropy. Since the crystal structure of MnxCoCrFeNi HEAs had FCC cubic structure, the values of BV and BR were equal. The calculated bulk modulus anisotropy index is zero, implying that the bulk modulus of the HEAs is isotropic. The variation trend of AG and AU values with Mn content is revealed in Figure 5d. The trend of AG and AU is consistent. The AG and AU of the CoCrFeNi HEA are the largest, which are 15.589 and 1.847, respectively. With the increase in Mn content, their values decreased slightly. When the value of x reaches 0.7, the values of AG and AU decrease to the lowest values of 7.96 and 0.853. As the x value continues to increase, the values of AG and AU increase slightly, but the change is not significant.
In order to more intuitively display the anisotropy of the alloy’s elastic modulus, the bulk modulus, shear modulus, and Young’s modulus of MnxCoCrFeNi HEAs are drawn in different directions using the spherical coordinate method. The directional relation of each cubic crystal modulus can be determined via the following equations [48]:
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 .
where Sij denotes the elastic compliance constants; l1, l2, and l3 denote the directional cosines; and l 1 = sin θ cos ϕ ;   l 2 = sin θ sin ϕ ;   l 3 = cos θ . When the 3D diagram is spherical, this means that the alloy is isotropic. Conversely, when the 3D diagram deviates from the sphere, the alloy is anisotropic. The greater the deviation, the greater the anisotropy. As illustrated in Figure 6, the anisotropic 3D diagrams of the bulk modulus, shear modulus and Young’s modulus of CoCrFeNi, Mn0.2CoCrFeNi, Mn0.5CoCrFeNi, Mn0.7CoCrFeNi, and MnCoCrFeNi are drawn, respectively. In Figure 6, the 3D figure of the bulk modulus is single in color and spherical, demonstrating that the bulk moduli of these HEAs are isotropic. The shear modulus and Young’s modulus are all deviated from the sphere, meaning that the shear modulus and Young’s modulus are anisotropic. The 3D diagram of the shear modulus of CoCrFeNi deviates most from the sphere, which clearly indicates that the shear modulus anisotropy of CoCrFeNi is the most severe. On the contrary, the deviation of the 3D diagram of the shear modulus of Mn0.7CoCrFeNi from the sphere is relatively small, so the shear modulus anisotropy of Mn0.7CoCrFeNi is small. The 3D diagram of the Young’s modulus indicates a similar trend to that of the shear modulus. These results are consistent with the previous analysis results of AB and AG.

3.3. Electronic Properties

In order to study the electronic properties of MnxCoCrFeNi high-entropy alloys with different Mn concentrations, the density of states and charge density difference of CoCrFeNi, Mn0.5CoCrFeNi, and MnCoCrFeNi are also calculated. The total and partial density of states is plotted from −15 to 40 eV. The Fermi energy level (Ef) at 0 eV is marked with a dashed line. As can be seen from Figure 7, the d-orbital electrons have a strong value compared with the s-orbital and p-orbital electrons in the whole range. The TDOS is mainly contributed by the d-orbital, and the peak of the TDOS is mainly distributed in the range of −5–2 eV near the Fermi level. In the case of x = 0, 0.5, and 1.0, the total densities of states at Fermi level are 45.05, 120.44, and 124.69, respectively. Obviously, the total density of states increases as the value of x increases. It can be inferred that the stability of the HEAs decrease with the increase in Mn concentration.
In the VCA model, all atoms in a cell are the same, so the calculation of charge density difference is related to the atomic density difference in FCC cell and the VCA model. In the end, the differences in electron density of MnxCoCrFeNi HEAs are simply compared. Figure 8 shows a two-dimensional slice of the charge density difference of (110) plane in MnxCoCrFeNi (x = 0, 0.5, and 1) HEAs. The plotting range of charge density difference is from −0.1 to 0.1 e/Å3. Electron loss is represented in blue and electron accumulation in red. In Figure 8, is evident that with the increase in Mn content, the degree of charge transfer gradually decreases. The weakening of atomic bonding leads to the instability of MnxCoCrFeNi, which can increase the possibility of dislocation.

3.4. Work Function

Generally, the electron work function (EWF) is the minimum amount of energy required for an electron in a metal to move to its surface at the Fermi level. It is one of the basic electronic properties of metals, depending on their composition and surface conditions. It fundamentally reflects the interaction between atoms and is directly related to the physical properties. In order to further understand the properties of MnxCoCrFeNi HEAs, the degrees of EWF of low exponent surfaces (100), (110), and (111) were calculated. The calculated trend of the EWF of MnxCoCrFeNi HEAs with a change in Mn concentration is drawn in Figure 9. It can be seen that the EWF of surface (110) is the largest, followed by surface (100) and surface (111). With the increase in Mn concentration, the EWF of (111), (110), and (100) decreases continuously. As the EWF decreases, the atomic bonds weaken. Therefore, with the raise of Mn concentration, dislocations are more likely to occur, which is in accordance with the results of previous GSFE analysis. The research results of Shi Y.Z. et al. [15] proved that with the decrease in the work function, the alloy system containing high passivation elements (Cr, Ni, Mo, Ti, etc.) has higher corrosion resistance. There were passivation elements Cr and Ni in MnxCoCrFeNi HEAs, and the EWF gradually decreased. It can be inferred that the corrosion resistance of MnxCoCrFeNi HEAs heightens when accompanied by Mn concentration. The study of Lu H. et al. [49] demonstrated that the Young’s modulus and hardness would increase with the increase in EWF and the strengthening of atomic bonds. It is worth noting that the work functions of MnxCoCrFeNi HEAs have different trends with respect to the Young’s modulus and hardness.

4. Conclusions

In this paper, the stacking fault energy, elastic properties, electronic properties, and work function of the MnxCoCrFeNi high-entropy alloy are analyzed using density functional theory. The calculated results can be summarized as follows.
The results of the GSFE curves demonstrate that the alloy with higher Mn content has lower anti-dislocation nucleation ability and easily forms dislocation. The relative barrier heights ( δ u s f u t f ) of the HEAs are all positive, which indicates that the deformation mechanism of the HEAs is not conducive to the transformation from dislocation to twin. The results of the GSFE and surface energies suggest that plastic deformation is controlled by dislocation-mediated slip. However, with the increase in Mn concentration, the value of δ u s f u t f decreases, which suggests that mechanical twinning can still be formed in the alloy. Based on the Rice-criterion, it is inferred that the MnxCoCrFeNi HEAs are ductile and have the formation of crack tip dislocation.
With the increase in Mn concentration, the bulk modulus, shear modulus, and Young’s modulus of MnxCoCrFeNi HEAs tend to increase. It can be concluded that with the enhancement of Mn concentration, the deformation resistance, shear resistance, and stiffness of the HEAs are improved. The density of states of MnxCoCrFeNi (x = 0, 0.5, and 1) HEAs at the Fermi level decreases with the increase in x, which indicates that the stability of the alloy decreases. The calculation of charge density difference demonstrates that the degree of charge transfer decreases gradually. This phenomenon will weaken the atomic bond and lead to an increase in the instability of the alloy, thus increasing the possibility of dislocation. However, the work function variation trend of MnxCoCrFeNi HEAs with different Mn concentrations is different from the Young’s modulus.

Author Contributions

Methodology, D.L.; Software, F.S. and Y.F.; Validation, D.L.; Formal analysis, H.X.; Investigation, G.Z.; Writing—original draft, G.Z.; Writing—review & editing, H.X.; Visualization, G.Z.; Funding acquisition, F.S. All authors have read and agreed to the published version of the manuscript.

Funding

The present study was financially supported by the Fund for Shanxi “1331 Project”, the Startup Fund for Talent Introduction of Shanxi Electronic Science and Technology Institute “2023RKJ029”.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yeh, J.W.; Chen, S.K.; Lin, S.J.; Gan, J.Y.; Chin, T.S.; Shun, T.T.; Tsau, C.H.; Chan, S.Y. Nanostructured high-entropy alloys with multiple principal elements: Novel alloy design concepts and outcomes. Adv. Eng. Mater. 2004, 6, 299–303. [Google Scholar] [CrossRef]
  2. Liu, W.H.; Lu, Z.P.; He, J.Y.; Luan, J.H.; Wang, Z.J.; Liu, B.; Liu, Y.; Chen, M.W.; Liu, C.T. Ductile CoCrFeNiMox high entropy alloys strengthened by hard intermetallic phases. Acta Mater. 2016, 116, 332–342. [Google Scholar] [CrossRef]
  3. Tao, L.; Sun, M.; Zhou, Y.; Luo, M.; Lv, F.; Li, M.; Zhang, Q.; Gu, L.; Huang, B.; Guo, S. A general synthetic method for high-entropy alloy subnanometer ribbons. J. Am. Chem. Soc. 2022, 144, 10582–10590. [Google Scholar] [CrossRef]
  4. Tsai, K.Y.; Tsai, M.H.; Yeh, J.W. Sluggish diffusion in Co-Cr-Fe-Mn-Ni high-entropy alloys. Acta Mater. 2013, 61, 4887–4897. [Google Scholar] [CrossRef]
  5. Cao, B.; Wang, C.; Yang, T.; Liu, C. Cocktail effects in understanding the stability and properties of face-centered-cubic high-entropy alloys at ambient and cryogenic temperatures. Scr. Mater. 2020, 187, 250–255. [Google Scholar] [CrossRef]
  6. Li, J.; Chen, H.; Fang, Q.; Jiang, C.; Liu, Y.; Liaw, P.K. Unraveling the dislocation–precipitate interactions in high-entropy alloys. Int. J. Plast. 2020, 133, 102819. [Google Scholar] [CrossRef]
  7. Ding, Q.; Zhang, Y.; Chen, X.; Fu, X.; Chen, D.; Chen, S.; Gu, L.; Wei, F.; Bei, H.; Gao, Y.; et al. Tuning element distribution, structure and properties by composition in high-entropy alloys. Nature 2019, 574, 223–227. [Google Scholar] [CrossRef]
  8. Ma, E. Unusual dislocation behavior in high-entropy alloys. Scr. Mater. 2020, 181, 127–133. [Google Scholar] [CrossRef]
  9. Qin, G.; Chen, R.; Liaw, P.K.; Gao, Y.; Li, X.; Zheng, H.; Wang, L.; Su, Y.; Guo, J.; Fu, H. A novel face-centered-cubic high-entropy alloy strengthened by nanoscale precipitates. Scr. Mater. 2019, 172, 51–55. [Google Scholar] [CrossRef]
  10. Yan, X.H.; Zhang, Y. Functional properties and promising applications of high entropy alloys. Scr. Mater. 2020, 187, 188–193. [Google Scholar] [CrossRef]
  11. Cantor, B.; Chang, I.T.H.; Knight, P.; Vincent, A.J.B. Microstructural development in equiatomic multicomponent alloys. Mater. Sci. Eng. A 2004, 375, 213–218. [Google Scholar] [CrossRef]
  12. Li, Z.; Pradeep, K.G.; Deng, Y.; Raabe, D.; Tasan, C.C. Metastable high-entropy dual-phase alloys overcome the strength-ductility trade-off. Nature 2016, 534, 227–230. [Google Scholar] [CrossRef] [PubMed]
  13. Sun, X.; Zhang, H.; Lu, S.; Ding, X.D.; Wang, Y.Z.; Vitos, L. Phase selection rule for Al-doped CrMnFeCoNi high-entropy alloys from first-principles. Acta Mater. 2017, 140, 366–374. [Google Scholar] [CrossRef]
  14. Zhang, H.; Sun, X.; Lu, S.; Dong, Z.; Ding, X.; Wang, Y.; Vitos, L. Elastic properties of AlxCrMnFeCoNi (0 ≤ x ≤ 5) high-entropy alloys from ab initio theory. Acta Mater. 2018, 155, 12–22. [Google Scholar] [CrossRef]
  15. Shi, Y.Z.; Collins, L.; Feng, R.; Zhang, C.; Balke, N.; Liaw, P.K.; Yang, B. Homogenization of AlxCoCrFeNi high-entropy alloys with improved corrosion resistance. Corros. Sci. 2018, 133, 120–131. [Google Scholar] [CrossRef]
  16. Zhang, J.; Xiong, K.; Huang, L.; Xie, B.; Ren, D.; Tang, C.; Feng, W. Effect of Doping with Different Nb Contents on the Properties of CoCrFeNi High-Entropy Alloys. Materials 2023, 16, 6407. [Google Scholar] [CrossRef]
  17. Kivy, M.B.; Zaeem, M.A. Generalized stacking fault energies, ductilities, and twinnabilities of CoCrFeNi-based face-centered cubic high entropy alloys. Scr. Mater. 2017, 139, 83–86. [Google Scholar] [CrossRef]
  18. Achmad, T.L.; Fu, W.; Chen, H.; Zhang, C.; Yang, Z.G. First-principles calculations of generalized-stacking-fault-energy of Co-based alloys. Comput. Mater. Sci. 2016, 121, 86–96. [Google Scholar] [CrossRef]
  19. Liu, S.F.; Wu, Y.; Wang, H.T.; He, J.Y.; Liu, J.B.; Chen, C.X.; Liu, X.J.; Wang, H.; Lu, Z.P. Stacking fault energy of face-centered-cubic high entropy alloys. Intermetallics 2018, 93, 269–273. [Google Scholar] [CrossRef]
  20. Lu, J.; Hultman, L.; Holmström, E.; Antonsson, K.H.; Grehk, M.; Li, W.; Vitos, L.; Golpayegani, A. Stacking fault energies in austenitic stainless steels. Acta Mater. 2016, 111, 39–46. [Google Scholar] [CrossRef]
  21. Xu, X.; Liu, P.; Tang, Z.; Hirata, A.; Song, S.; Nieh, T.; Liaw, P.; Liu, C.; Chen, M. Transmission electron microscopy characterization of dislocation structure in a face-centered cubic high-entropy alloy Al0.1CoCrFeNi. Acta Mater. 2018, 144, 107–115. [Google Scholar] [CrossRef]
  22. Ullrich, C.; Eckner, R.; Krüger, L.; Martin, S.; Klemm, V.; Rafaja, D. Interplay of microstructure defects in austenitic steel with medium stacking fault energy. Mater. Sci. Eng. A 2016, 649, 390–399. [Google Scholar] [CrossRef]
  23. Zhang, Y.; Guo, J.; Chen, J.; Wu, C.; Kormout, K.S.; Ghosh, P.; Zhang, Z. On the stacking fault energy related deformation mechanism of nanocrystalline Cu and Cu alloys: A first-principles and TEM study. J. Alloys Compd. 2019, 776, 807–818. [Google Scholar] [CrossRef]
  24. Mayahi, R. An investigation concerning generalized stacking fault behavior of AlCoxCrFeNi (0.25 ≤ x ≤ 2) high entropy alloys: Insights from first-principles study. J. Alloys Compd. 2020, 818, 152928. [Google Scholar] [CrossRef]
  25. Zhao, Q.; Li, J.; Fang, Q.; Feng, H. Effect of Al solute concentration on mechanical properties of AlxFeCuCrNi high entropy alloys: A first-principles study. Phys. B 2019, 566, 30–37. [Google Scholar] [CrossRef]
  26. Nong, Z.S.; Lei, Y.N.; Zhu, J.C. Wear and oxidation resistances of AlCrFeNiTi-based high entropy alloys. Intermetallics 2018, 101, 144–151. [Google Scholar] [CrossRef]
  27. Pan, Y.; Pu, D. First-principles investigation of oxidation behavior of Mo5SiB2. Ceram. Int. 2020, 46, 6698–6702. [Google Scholar] [CrossRef]
  28. Mitro, S.K.; Hossain, K.M.; Majumder, R.; Hasan, M.Z. Effect of the negative chemical pressure on physical properties of doped perovskite molybdates in the framework of DFT method. J. Alloys Compd. 2021, 854, 157088. [Google Scholar] [CrossRef]
  29. Vu, V.T.; Lavrentyev, A.A.; Gabrelian, B.V.; Parasyuk, O.V.; Ocheretova, V.A.; Khyzhun, O.Y. Electronic structure and optical properties of Ag2HgSnSe4: First-principles DFT calculations and X-ray spectroscopy studies. J. Alloys Compd. 2018, 732, 372–384. [Google Scholar] [CrossRef]
  30. Peivaste, I.; Alahyarizadeh, G.; Minuchehr, A.; Aghaie, M. Comparative study on mechanical properties of three different SiCpolytypes (3C, 4H and 6H) under high pressure: First-principle calculations. Vacuum 2018, 154, 37–43. [Google Scholar] [CrossRef]
  31. Chen, J.Y.; Zhang, X.D.; Ying, C.H.; Ma, H.; Li, J.; Wang, F.; Guo, H. The influence of vacancy defects on elastic and electronic properties of TaSi (5/3) desilicides from a first-principles calculations. Ceram. Int. 2020, 46, 10992–10999. [Google Scholar] [CrossRef]
  32. Pang, X.Z.; Yang, W.C.; Yang, J.B.; Pang, M.J.; Zhan, Y.Z. Atomic structure, stability and electronic properties of S(Al2CuMg)/Al interface: A first-principles study. Intermetallics 2018, 93, 329–337. [Google Scholar] [CrossRef]
  33. Chen, S.; Pan, Y.; Wang, D.; Deng, H. Structural stability and electronic and optical properties of bulk WS2 from first-principles investigations. Mat. Sci. Eng. B-Solid. 2020, 259, 114580. [Google Scholar] [CrossRef]
  34. Zaddach, A.J.; Niu, C.; Koch, C.C.; Irving, D.L. Mechanical properties and stacking fault energies of NiFeCrCoMn high-entropy alloy. Jom 2013, 65, 1780–1789. [Google Scholar] [CrossRef]
  35. Li, X.; Irving, D.L.; Vitos, L. First-principles investigation of the micromechanical properties of fcc-hcp polymorphic high-entropy alloys. Sci. Rep. 2018, 8, 11196. [Google Scholar] [CrossRef]
  36. Cantor, B. Multicomponent and high entropy alloys. Entropy 2014, 16, 4749–4768. [Google Scholar] [CrossRef]
  37. Udagawa, Y.; Yamaguchi, M.; Abe, H.; Sekimura, N.; Fuketa, T. Ab initio study on plane defects in zirconium-hydrogen solid solution and zirconium hydride. Acta Mater. 2010, 58, 3927–3938. [Google Scholar] [CrossRef]
  38. Swygenhoven, H.V.; Derlet, P.M.; Frøseth, A.G. Stacking fault energies and slip in nanocrystalline metals. Nat. Mater. 2004, 3, 399–403. [Google Scholar] [CrossRef]
  39. Tadmor, E.B.; Hai, S. A Peierls criterion for the onset of deformation twinning at a crack tip. J. Mech. Phys. Solids 2003, 51, 765–779. [Google Scholar] [CrossRef]
  40. Asaro, R.J.; Suresh, S. Mechanistic models for the activation volume and rate sensitivity in metals with nanocrystalline grains and nano-scale twins. Acta Mater. 2005, 53, 3369–3382. [Google Scholar] [CrossRef]
  41. Kibey, S.; Liu, J.B.; Johnson, D.D.; Sehitoglu, H. Predicting twinning stress in fcc metals: Linking twin-energy pathways to twin nucleation. Acta Mater. 2007, 55, 6843–6851. [Google Scholar] [CrossRef]
  42. Rice, J.R. Dislocation nucleation from a crack tip an analysis based on the Peierls concept. J. Mech. Phys. Solids 1992, 40, 239–271. [Google Scholar] [CrossRef]
  43. Wen, Z.Q.; Zhao, Y.H.; Hou, H.; Wang, B.; Han, P.D. The mechanical and thermodynamic properties of Heusler compounds Ni2XAl (X = Sc, Ti, V) under pressure and temperature: A first-principles study. Mater. Des. 2017, 114, 398–403. [Google Scholar] [CrossRef]
  44. Sun, F.; Zhang, G.; Liu, H.; Xu, H.; Fu, Y.; Li, D. Effect of transition-elements substitution on mechanical properties and electronic structures of B2-AlCu compounds. Results Phys. 2021, 21, 103765. [Google Scholar] [CrossRef]
  45. Chen, X.Q.; Niu, H.; Li, D.; Li, Y. Modeling hardness of polycrystalline materials and bulk metallic glasses. Intermetallics 2011, 19, 1275. [Google Scholar] [CrossRef]
  46. Pugh, S.F. XCII. Relations between the elastic moduli and the plastic properties of polycrystalline pure metals. Philos. Mag. 1954, 45, 823–843. [Google Scholar] [CrossRef]
  47. Sun, F.; Zhang, G.; Ren, X.; Wang, M.; Xu, H.; Fu, Y.; Tang, Y.; Li, D. First-principles studies on phase stability, anisotropic elastic and electronic properties of Al-La binary system intermetallic compounds. Mater. Today Commun. 2020, 24, 101101. [Google Scholar] [CrossRef]
  48. Nye, J.F. Physical Properties of Crystals: Their Representation by Tensors and Matrices; Oxford University Press: Oxford, UK, 1985. [Google Scholar]
  49. Lu, H.; Liu, Z.; Yan, X.; Li, D.; Parent, L.; Tian, H. Electron work function-a promising guiding parameter for material design. Sci. Rep. 2016, 6, 24366. [Google Scholar] [CrossRef]
Disclaimer/Publisher’s Note: The statements, opinions, and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions, or products referred to in the content.
Figure 1. The computational cell used in the present work: (a) the atomic configuration of the FCC cell with stacking sequence A, B, and C; (b) the supercell consisting of 13 atomic layers of (111) plane.
Figure 1. The computational cell used in the present work: (a) the atomic configuration of the FCC cell with stacking sequence A, B, and C; (b) the supercell consisting of 13 atomic layers of (111) plane.
Materials 17 04378 g001
Figure 2. The perfect slab of (100), (110), and (111) planes containing 10 atomic layers with 5 Å vacuum layers in the bottom and top (a,c,e). Surface models of (100), (110), and (111) planes containing an 8 Å vacuum block in the center of the cell (b,d,f).
Figure 2. The perfect slab of (100), (110), and (111) planes containing 10 atomic layers with 5 Å vacuum layers in the bottom and top (a,c,e). Surface models of (100), (110), and (111) planes containing an 8 Å vacuum block in the center of the cell (b,d,f).
Materials 17 04378 g002
Figure 3. The calculated GSFE curves for MnxCoCrFeNi HEAs.
Figure 3. The calculated GSFE curves for MnxCoCrFeNi HEAs.
Materials 17 04378 g003
Figure 4. The calculated ratios of γisfusf (a), twinnability for crack tip twinning τa (b), twinnability for grain boundary twinning τ (c), anticipated critical stress τcrit (d), the Rice-criterion for ductility D (e), and the surface energies γs (f) for MnxCoCrFeNi HEAs.
Figure 4. The calculated ratios of γisfusf (a), twinnability for crack tip twinning τa (b), twinnability for grain boundary twinning τ (c), anticipated critical stress τcrit (d), the Rice-criterion for ductility D (e), and the surface energies γs (f) for MnxCoCrFeNi HEAs.
Materials 17 04378 g004
Figure 5. The effect of Mn concentration on MnxCoCrFeNi HEAs elastic properties: (a) elastic constants; (b) bulk modulus, shear modulus, Young’s modulus, and Vickers hardness; (c) Poisson’s ratio and G/B; (d) anisotropy indexes AG and AU.
Figure 5. The effect of Mn concentration on MnxCoCrFeNi HEAs elastic properties: (a) elastic constants; (b) bulk modulus, shear modulus, Young’s modulus, and Vickers hardness; (c) Poisson’s ratio and G/B; (d) anisotropy indexes AG and AU.
Materials 17 04378 g005
Figure 6. The anisotropic characteristics of the bulk modulus, shear modulus, and Young’s modulus of MnxCoCrFeNi HEAs.
Figure 6. The anisotropic characteristics of the bulk modulus, shear modulus, and Young’s modulus of MnxCoCrFeNi HEAs.
Materials 17 04378 g006
Figure 7. Total and partial density of MnxCoCrFeNi HEAs (x = 0, 0.5, and 1).
Figure 7. Total and partial density of MnxCoCrFeNi HEAs (x = 0, 0.5, and 1).
Materials 17 04378 g007
Figure 8. The electron density difference of the (110) plane in MnxCoCrFeNi HEAs (x = 0, 0.5, and 1).
Figure 8. The electron density difference of the (110) plane in MnxCoCrFeNi HEAs (x = 0, 0.5, and 1).
Materials 17 04378 g008
Figure 9. The effect of Mn concentration on work functions of MnxCoCrFeNi HEAs (100), (110), and (111) planes.
Figure 9. The effect of Mn concentration on work functions of MnxCoCrFeNi HEAs (100), (110), and (111) planes.
Materials 17 04378 g009
Table 1. Chemical compositions (in at.%) of MnxCoCrFeNi HEAs.
Table 1. Chemical compositions (in at.%) of MnxCoCrFeNi HEAs.
HEAsMn (at.%)Co (at.%)Cr (at.%)Fe (at.%)Ni (at.%)
CoCrFeNi025.025.025.025.0
Mn0.2CoCrFeNi2.424.424.424.424.4
Mn0.4CoCrFeNi9.222.722.722.722.7
Mn0.6CoCrFeNi13.221.721.721.721.7
Mn0.8CoCrFeNi16.820.820.820.820.8
MnCoCrFeNi20.020.020.020.020.0
Table 2. Calculated lattice parameter (in Å) of HEAs in the VCA model compared with previous experiment (Expt.).
Table 2. Calculated lattice parameter (in Å) of HEAs in the VCA model compared with previous experiment (Expt.).
HEAsa a S Q S a C P A a E x p t .
CoCrFeNi3.4153.540 [34]3.529 [35]3.575 [34]
MnCoCrFeNi3.5173.540 [34]3.529 [35]3.597 [34]
3.59 [36]
Table 3. The calculated unstable γusf (mJ/m2), stable γisf (mJ/m2), unstable twinning γutf (mJ/m2), twinning 2γtsf (mJ/m2), energy barrier height δ u s f u t f , γisfusf, γutfusf parameter, Rice-criterion D and surface energies of (100), (110) and (111) planes (J/m2) in MnxCoCrFeNi HEAs.
Table 3. The calculated unstable γusf (mJ/m2), stable γisf (mJ/m2), unstable twinning γutf (mJ/m2), twinning 2γtsf (mJ/m2), energy barrier height δ u s f u t f , γisfusf, γutfusf parameter, Rice-criterion D and surface energies of (100), (110) and (111) planes (J/m2) in MnxCoCrFeNi HEAs.
HEAs(x)γusfγisfγutftsf δ u s f u t f γisfusfγutfusfDγs(111)γs(110)γs(100)
0.0808.84578.93935.74574.71126.900.7161.1571.1133.0023.4653.150
0.2722.68511.42829.74490.12107.060.7081.1481.1232.7063.2922.955
0.4622.57416.32720.54389.0497.960.6691.1571.2022.4953.1872.807
0.6521.78319.25613.32288.6091.550.6121.1751.2292.1372.9092.576
0.8410.07209.17491.73169.9781.660.5101.1991.4231.9452.7372.307
1.0276.17110.04340.2160.3664.050.3981.2321.6891.5552.3632.080
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Sun, F.; Zhang, G.; Xu, H.; Li, D.; Fu, Y. Theoretical Analysis of Stacking Fault Energy, Elastic Properties, Electronic Properties, and Work Function of MnxCoCrFeNi High-Entropy Alloy. Materials 2024, 17, 4378. https://doi.org/10.3390/ma17174378

AMA Style

Sun F, Zhang G, Xu H, Li D, Fu Y. Theoretical Analysis of Stacking Fault Energy, Elastic Properties, Electronic Properties, and Work Function of MnxCoCrFeNi High-Entropy Alloy. Materials. 2024; 17(17):4378. https://doi.org/10.3390/ma17174378

Chicago/Turabian Style

Sun, Fenger, Guowei Zhang, Hong Xu, Dongyang Li, and Yizheng Fu. 2024. "Theoretical Analysis of Stacking Fault Energy, Elastic Properties, Electronic Properties, and Work Function of MnxCoCrFeNi High-Entropy Alloy" Materials 17, no. 17: 4378. https://doi.org/10.3390/ma17174378

APA Style

Sun, F., Zhang, G., Xu, H., Li, D., & Fu, Y. (2024). Theoretical Analysis of Stacking Fault Energy, Elastic Properties, Electronic Properties, and Work Function of MnxCoCrFeNi High-Entropy Alloy. Materials, 17(17), 4378. https://doi.org/10.3390/ma17174378

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