Next Article in Journal
Effect of Deformation-Induced Plasticity in Low-Alloyed Al-Mg-Zr Alloy Processed by High-Pressure Torsion
Next Article in Special Issue
Investigation of the Quenching Sensitivity of the Mechanical and Corrosion Properties of 7475 Aluminum Alloy
Previous Article in Journal
A Graphical Tool to Describe the Operating Point of Direct Reduction Shaft Processes
Previous Article in Special Issue
A New Method for Preparing Titanium Aluminium Alloy Powder
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

First-Principle Investigation into Mechanical Properties of Al6Mg1Zr1 under Uniaxial Tension Strain on the Basis of Density Functional Theory

1
School of Science, Shanghai Maritime University, Shanghai 201306, China
2
School of Mechanical and Energy Engineering, Shanghai Technical Institute of Electronics and Information, Shanghai 201411, China
3
College of Science, Inner Mongolia University of Technology, Hohhot 010051, China
*
Authors to whom correspondence should be addressed.
First authors.
Metals 2023, 13(9), 1569; https://doi.org/10.3390/met13091569
Submission received: 5 July 2023 / Revised: 14 August 2023 / Accepted: 29 August 2023 / Published: 7 September 2023
(This article belongs to the Special Issue Aluminum Alloys and Aluminum-Based Matrix Composites)

Abstract

:
The influences of uniaxial tension strain in the x direction (εx) on the mechanical stability, stress–strain relations, elastic properties, hardness, ductility, and elastic anisotropy of Al6Mg1Zr1 compound were studied by performing first-principle calculations on the basis of density functional theory. It was found that Al6Mg1Zr1 compound is mechanically stable in the range of strain εx from 0 to 6%. As the strain εx increased from 0 to 6%, the stress in the x direction (σx) first grew linearly and then followed a nonlinear trend, while the stresses in the y and z directions (σy and σz) showed a linearly, increasing trend all the way. The bulk modulus B, shear modulus G, and Young’s modulus E all dropped as the strain εx increased from 0 to 6%. The Poisson ratio μ of Al6Mg1Zr1 compound was nearly unchanged when the strain εx was less than 3%, but then it grew quickly. Vickers hardness HV of Al6Mg1Zr1 compound dropped gradually as the strain εx increased from 0 to 6%. The Al6Mg1Zr1 compound was brittle when the εx was less than 4%, but it presented ductility when the strain εx was more than 4%. As the strain εx increased from 0 to 6%, the compression anisotropy percentage (AB) grew and its slope became larger when the strain εx was more than 4%, while both the shear anisotropy percentage (AG) and the universal anisotropy index (AU) first dropped slowly and then grew quickly. These results demonstrate that imposing appropriate uniaxial tension strain can affect and regulate the mechanical properties of Al6Mg1Zr1 compound.

1. Introduction

From an industrial point of view, the aluminum–magnesium (Al-Mg) based alloys and intermetallic compounds are considered as very promising engineering materials, and have been widely used in aviation, aerospace, shipbuilding, rail transportation, and automotive industries owing to their high strength-to-weight ratio, good formability, excellent corrosion resistance, and good weldability [1,2,3,4]. However, Al-Mg based materials only have a low to medium strength, which severely restricts their application in industry [5,6,7]. With the rapid development of industrial technology, a further improvement of the comprehensive performance of Al-Mg based materials is required. An effective approach to improve the performance of the Al-Mg based materials is by adding the appropriate elements [8,9,10,11,12]. The rare-earth metal scandium (Sc) has proved to be the most effective element to improve the performance of the Al-Mg based materials [13,14,15,16]. However, the application of Sc is greatly limited in industry due to its high cost. Therefore, it is of great significance to search for another lower-cost additional element to replace Sc. The transition metal zirconium (Zr) is of much lower cost (its price is only about 1/100 of that of Sc), and can serve a similar strengthening function to Sc in Al-Mg based materials [17,18,19,20,21].
In recent years, it was reported that the functional properties of materials were regulated by imposing strain [22,23,24,25]. Fu, et al. quantitatively analyzed the effects of heterogeneous plastic strain on hydrogen-induced cracking of twinning-induced plasticity (TWIP) steel by electron backscattered diffraction (EBSD) technology. According to a quantitative calculation of EBSD crystallographic data, the larger the geometrically necessary dislocation density (ρGND), the greater is the heterogeneous strain. It was evident from the overall statistical analyses that hydrogen-induced crack initiation depends on interactions between heterogeneous strain and hydrogen atoms generated by local grain orientation deviation (deviation from ideal) and gradient (misorientation). The larger the ρGND, the easier it is to cause local enrichment of hydrogen atoms, which in turn leads to hydrogen-induced cracking in TWIP steel [26]. Liang et al. investigated the strain-induced strengthening in superconducting β-Mo2C through high pressure and high temperature. It was found that strain-induced high-density dislocations and low-angle grain boundaries were introduced and enabled the synthesized β-Mo2C ceramics to exhibit surprising mechanical properties [27]. Du, et al. studied the Poisson ratio of in-plane pristine armchair and zigzag graphene under uniaxial tensile loading by molecular dynamics simulations, which indicated that the Poisson ratio strongly depends on the tensile strain. At the critical strain, the Poisson ratio will transform from positive to negative, and the critical strain of the zigzag is far less than that of armchair [28]. Rasidul Islam, et al. investigated the strain-induced mechanical properties of inorganic halide perovskite CsGeBr3 through first-principles based on density functional theory. It was found that the bulk modulus, shear modulus, and Young’s modulus all increased with increasing compressive strain but decreased with increasing tensile strain. The brittleness of CsGeBr3 increased with compressive strain, whereas CsGeBr3 offered significant ductility with more than 2% tensile strain [29]. Tan et al. investigated the mechanical behavior of AlSi2Sc2 under uniaxial tensile strain by performing first-principle calculations based on density functional theory. It was found that the estimated elastic moduli of AlSi2Sc2 decreased with increasing uniaxial tensile strain, but the brittleness of AlSi2Sc2 did not change when strain was applied [30]. However, there has been little investigation into the effect of strain on the mechanical properties of Al-Mg-Zr compounds based on the density functional theory.
In the present study, the first principle calculations based on density functional theory were used to investigate the elastic properties, hardness, ductility, and elastic anisotropy of Al6Mg1Zr1 compound at different uniaxial tensile strains. The current research will contribute to a better understanding of the modulation of strain on the mechanical properties of Al-Mg based alloys and compounds.

2. Computational Methods

In this study, the structural model of Al6Mg1Zr1 supercell viewed along the c axis is shown in Figure 1. Grey, orange, and green spheres represent Al, Mg, and Zr atoms, respectively. The x, y, and z direction are parallel to the a, b, and c axis, respectively. Our calculations of Al6Mg1Zr1 were carried out by the Cambridge Serial Total Energy Package (CASTEP) code, using the plane-wave pseudopotential method based on density functional theory (DFT) [31,32,33]. For the exchange and correlation terms in the electron–electron interaction, the generalized gradient approximation (GGA) in the scheme of Perdew–Burke–Ernzerhof (PBE) was used [34]. The valence wave functions were expanded in a plane-wave basis set up to an energy cutoff of 600 eV. For the k point sampling, a 3 × 6 × 6 Monkhorst–Pack mesh in the Brillouin zone was used. The other parameters used default settings of ultra-fine accuracy.

3. Results and Discussion

3.1. Mechanical Stability

Elastic stiffness constants Cij are very important physical quantities used to express the elasticity of a solid material in engineering applications. In this work, the elastic stiffness constants were calculated using the stress–strain approach based on Hook’s law by imposing uniaxial tension strain in the x direction (εx). The elastic stiffness constants Cij of Al6Mg1Zr1 compound at different εx are presented in Table 1. It can be found that there are nine independent effective constants (C11, C12, C13, C22, C23, C33, C44, C55, and C66) in the elastic stiffness matrix for Al6Mg1Zr1. Therefore, Al6Mg1Zr1 is determined to be of orthorhombic structure [35].
Based on Born–Huang’s dynamical theory of crystal lattices, the mechanical stability standards for orthorhombic crystals must meet the following requirements [36]:
C i i > 0 C 11 + C 22 2 C 12 > 0 C 11 + C 33 2 C 13 > 0 C 22 + C 33 2 C 23 > 0 C 11 + C 22 + C 33 + 2 ( C 12 + C 13 + C 23 ) > 0
It was noted that the elastic constants Cij of the Al6Mg1Zr1 fulfilled well the mechanical stability standards in the range of strain εx from 0 to 6%, while they could not meet the standards when the εx was more than 6%. Therefore, the orthorhombic of Al6Mg1Zr1 is mechanically stable in the range of strain εx from 0 to 6%. In this study, we are only concerned with the mechanical properties of Al6Mg1Zr1 in the range of strain εx from 0 to 6%.

3.2. Stress–Strain Relations

The stresses in the principal axis direction (σx, σy, and σz) of Al6Mg1Zr1 at different uniaxial tension strains in the x direction (εx) were calculated based on Hook’s law, and the calculated values are presented in Table 2 and Figure 2.
It can be seen that, as the strain εx increased from 0 to 6%, the stress in the x direction (i.e., σx) gradually grew to 7.7129 GPa. The σx-εx curve of Al6Mg1Zr1 had a linear formation in the range of strain εx from 0 to 3%, and then followed a nonlinear trend in the range of strain εx from 3% to 6%. It is indicated that the deformation of Al6Mg1Zr1 was elastic in the range of εx from 0 to 3%, after which plastic deformation occurred. The stresses in the y and z directions (σy and σz) were almost equal and had good linear relation with the uniaxial tension strain in the range of εx from 0 to 6%. On the whole, the stress σx was much higher than σy and σz due to the uniaxial tension loading being in the x direction.

3.3. Elastic Properties of Polycrystalline Materials

In many cases, polycrystalline materials have advantages in practical applications compared to single crystal materials [37]. Therefore, it is more meaningful to examine the elastic properties of polycrystalline materials. The elastic properties of polycrystalline materials can be characterized by the bulk modulus B, shear modulus G, Young’s modulus E, and Poisson ratio μ.
There are two approximation methods to obtain polycrystalline elastic moduli, namely, the Voigt method and the Reuss method. The Voigt method provides the upper bound to the polycrystalline elastic moduli, and the Reuss method provides the lower bound to the polycrystalline elastic moduli. For different crystalline systems, the bulk modulus B and shear modulus G according to Voigt and Reuss approximations are given by the following equations [38]:
B V = C 11 + C 22 + C 33 + 2 ( C 12 + C 13 + C 23 ) 9
G V = C 11 + C 22 + C 33 C 12 C 13 C 23 15 + C 44 + C 55 + C 66 5
B R = 1 S 11 + S 22 + S 33 + 2 ( S 12 + S 13 + S 23 )
G R = 15 4 ( S 11 + S 22 + S 33 ) + 3 ( S 44 + S 55 + S 66 ) 4 ( S 12 + S 13 + S 23 )
where the subscripts V and R denote the Voigt and Reuss averages, Cij are the elastic stiffness constants, and Sij are the elastic compliance coefficients.
The arithmetic average of the Voigt and the Reuss bounds is referred to as the Voigt–Reuss Hill (VRH) average, and it is considered as the best estimate of the theoretical polycrystalline elastic moduli. The VRH averages of B and G are given as follows [38]:
B = B V + B R 2
G = G V + G R 2
The Young’s modulus E and Poisson ratio μ of the polycrystalline material can be obtained from the bulk modulus B and shear modulus G, while the corresponding calculation formulas are as follows [38]:
E = 9 B G 3 B + G
μ = 3 B 2 G 6 B + 2 G
The Bulk modulus B, shear modulus G, Young’s modulus E and Poisson ratio μ of polycrystalline Al6Mg1Zr1 at different uniaxial tension strains in the x direction (εx) were calculated using the above formulas, and the calculated values are presented in Table 3 and Figure 3.
The bulk modulus B is a measure of the resistance of a solid material to a volume change. Figure 3a presents the calculated bulk modulus B of Al6Mg1Zr1 at different strains εx. It can be seen that, as the strain εx increased from 0 to 6%, the bulk modulus B dropped from 82.93 GPa to 60.63 GPa. The bulk modulus B reduced by 26.9%, which showed its negative relation with the uniaxial tension strain for Al6Mg1Zr1. The Al6Mg1Zr1 alloy has the largest incompressibility at the unstrained state due to the largest B value, while it has the largest compressibility at the strain εx of 6% due to the smallest B value.
The shear modulus G is defined as the ratio of shear stress to the shear strain, which characterizes the ability of a solid material to resist deformation under shear stress. The greater G corresponds to the stronger shear resistance of the solid material. Figure 3b presents the calculated values of shear modulus G of Al6Mg1Zr1 at different strains εx. It can be seen that as the strain εx increased from 0 to 6%, the shear modulus G dropped from 53.77 GPa to 22.88 GPa. The shear modulus G was reduced by 57.5%, which indicated that shear resistance is greatly influenced by the uniaxial tension strain. Al6Mg1Zr1 alloy has the smallest shear resistance at the strain εx of 6% due to the smallest G value.
Young’s modulus E is defined as the ratio of tensile stress and axial strain and serves as a measure of the stiffness of solid materials. Figure 3c presents the calculated values of Young’s modulus E of Al6Mg1Zr1 at different strains εx. It can be found that Young’s modulus E dropped with increasing strain εx. When the Al6Mg1Zr1 was unstrained, the Young’s modulus E was 132.63 GPa. When the strain εx reached 6%, Young’s modulus E dropped to 60.96 GPa. Young’s modulus E was reduced by 54.0%, showing its negative relation with uniaxial tension strain. The unstrained Al6Mg1Zr1 has the largest stiffness due to the largest E value, while Al6Mg1Zr1 alloy has the smallest stiffness at the strain εx of 6% due to the smallest E value.
Figure 3d presents the calculated values of Poisson ratio μ of Al6Mg1Zr1 at different strains εx. It can be found that when the strain εx was less than 3%, the Poisson ratio μ remained nearly unchanged. When εx was more than 3%, the Poisson ratio μ grew quickly with increasing strain εx, showing its positive relation with uniaxial tension strain. Generally, when the Poisson ratio μ is between −1 and 0.5, the solid is relatively stable under shear deformation. From Figure 3d, it can be seen that the Poisson ratio μ of Al6Mg1Zr1 ranged from 0.23 to 0.33, which is between −1 and 0.5, indicating that Al6Mg1Zr1 is a stable linear elastic solid. Al6Mg1Zr1 has the maximum μ value at the strain εx of 6%, indicating that Al6Mg1Zr1 has the highest toughness at the strain εx of 6%.
By comparing Figure 3a–c and Figure 3d, it can be found that with the strain εx increasing from 0 to 6%, the elastic moduli (B, G, and E) monotonically decreased with the increasing strain εx, while the Poisson ratio μ was first nearly unchanged and then grew quickly. The variation trends of elastic moduli (B, G, and E) and Poisson ratio μ of Al6Mg1Zr1 alloy with the uniaxial tensile strain εx are similar to that of AlSi2Sc2 [30].

3.4. Hardness and Ductility

Hardness is a measure of the resistance to localized deformation induced by either mechanical indentation or abrasion. In general, hardness is linked with the elastic and plastic properties of a material, and the shear modulus G is the more important parameter governing hardness than the bulk modulus B. There have been some semi-empirical models developed to predict the hardness of materials. Chen et al. [39] proposed a model to predict the Vickers hardness HV of polycrystalline materials and bulk metallic glasses based on the shear modulus G and the Pugh’s modulus ratio k (k = G/B) as follows [39]:
H V = 1.887 k 1.171 G 0.591
The above formula has been successfully used to predict the hardness of many compounds.
The ductility or brittleness of a solid material can be estimated by the ratio of the bulk modulus B to the elastic shear modulus G (i.e., B/G). If the modulus ratio B/G is more than 1.75, the solid is classified as ductile material; otherwise, it is brittle [40,41].
Table 4 and Figure 4 present the calculated values of Vickers hardness Hv and the modulus ratio B/G of Al6Mg1Zr1 at different uniaxial tension strains εx.
From Figure 4a, as the strain εx increased from 0 to 6%, the Vickers hardness HV of Al6Mg1Zr1 dropped gradually from 11.97 GPa to 3.83 Gpa. The Vickers hardness HV of Al6Mg1Zr1 dropped by 71.746%, showing its negative relation with the uniaxial tension strain.
From Figure 4b, when the strain εx was less than 3%, the modulus ratio B/G remained nearly unchanged. When the εx was more than 3%, the modulus ratio B/G grew quickly from 1.55 to 2.65 with the increasing strain εx. When the strain εx was less than 4%, the modulus ratio B/G < 1.75 and the Al6Mg1Zr1 was brittle. When the εx was more than 4%, the modulus ratio B/G > 1.75 and the Al6Mg1Zr1 was taken as ductile. Al6Mg1Zr1 would become most ductile at the strain of 6% due to the largest modulus ratio B/G value of 2.65.

3.5. Elastic Anisotropy

Elastic anisotropy of a solid material is very important in diverse applications such as phase transformations and dislocation dynamics, and it can be characterized by the elastic anisotropy indexes. The elastic anisotropy indexes include compression anisotropy percentage AB, shear anisotropy percentage AG, and the universal anisotropy index AU, and can be calculated using the following expressions [42]:
A B = B V B R B V + B R A G = G V G R G V + G R A U = 5 G V G R + B V B R 6
If AU = AB = AG = 0, the material shows characteristics of elastic isotropy. Otherwise, it has elastic anisotropy, and the larger the deviation of elastic anisotropy index values from 0 (the corresponding elastic isotropy value), the greater its degree in elastic anisotropy.
Table 5 and Figure 5 show the calculated values of elastic anisotropy indexes (AB, AG, and AU) of Al6Mg1Zr1 at different uniaxial tension strains εx.
From Figure 5a, the compression anisotropy percentage AB of Al6Mg1Zr1 grew slowly as the strain εx increased from 0 to 4%, and then it grew quickly as the strain εx increased from 4% to 6%. AB reached its maximum of 4.58% at the strain εx of 6%. From Figure 5b, when the strain εx was less than 3%, the shear anisotropy percentage AG dropped slowly with the strain εx. AG reached its minimum of 0.32% at the strain εx of 3%. When the strain εx was more than 3%, AG grew quickly with the strain εx, and reached its maximum of 7.37% at the strain εx of 6%. The universal anisotropy index AU can reflect the anisotropy more accurately because both bulk modulus B and shear modulus G are deliberated in its expression. As shown in Figure 5c, the change trend of AU with the strain εx is similar to that of AG. The variation trends of elastic anisotropy indexes of Al6Mg1Zr1 with uniaxial tensile strain are similar to that of hexagonal C40 MoSi2 [43]. Moreover, the values of the elastic anisotropy indexes (AB, AG, and AU) are small and the degree in elastic anisotropy of Al6Mg1Zr1 is relatively weak in the range of strain εx from 0 to 6%.

4. Conclusions

To sum up, in this work the mechanical stability, stress–strain relations, elastic properties, hardness, ductility, and elastic anisotropy of Al6Mg1Zr1 at different uniaxial tension strains in the x direction (εx) were examined by first principle calculations based on density functional theory. The influences of strain εx on the mechanical properties of the Al6Mg1Zr1 were studied. It was found that Al6Mg1Zr1 was mechanically stable in the range of strain εx from 0 to 6%, while it was unstable when the strain εx was more than 6%. As the strain εx increased from 0 to 6%, the stress in the x direction (σx) first increased linearly and then followed a nonlinear trend, but the stresses in the y and z directions (σy and σz), which were almost equal, showed a linear, increasing trend all the way. Due to the uniaxial tension loading in the x direction, the stress σx was much higher than σy and σz. The bulk modulus B, shear modulus G and Young’s modulus E of Al6Mg1Zr1 all dropped with increasing strain εx from 0 to 6%, showing their negative relations with uniaxial tension strain. Therefore, the incompressibility, shear resistance, and stiffness of the Al6Mg1Zr1 all dropped with increasing uniaxial tension strain. When the strain εx was less than 3%, The Poisson ratio μ of Al6Mg1Zr1 was nearly unchanged. However, it grew quickly when the εx was more than 3%, showing its positive relation with uniaxial tension strain. The Poisson ratio μ ranged from 0.23 to 0.33, which is between −1 and 0.5, indicating that Al6Mg1Zr1 is stable linear elastic solid in the range of εx from 0 to 6%. Al6Mg1Zr1 has the highest toughness at the strain εx of 6% due to the maximum μ value. The Vickers hardness HV of Al6Mg1Zr1 dropped gradually with the increasing strain εx from 0 to 6%, showing its negative relation with uniaxial tension strain. When the strain εx was less than 3%, the modulus ratio B/G of Al6Mg1Zr1 remained nearly unchanged but it grew quickly when the εx was more than 3%, showing its positive relation with uniaxial tension strain. Al6Mg1Zr1 was brittle when the εx was less than 4%, while it exhibited ductility when the strain εx was more than 4%. The best ductility was achieved for Al6Mg1Zr1 alloy at the strain εx of 6% due to the maximum B/G value. The compression anisotropy percentage AB of Al6Mg1Zr1 grew slowly as the strain εx increased from 0 to 4%, while it grew quickly as the strain εx increased from 4% to 6%. Both the shear anisotropy percentage (AG) and universal anisotropy index (AU) dropped slowly with increasing strain εx from 0 to 3%, and then grew quickly with increasing strain εx from 3 to 6%. In addition, the values of the elastic anisotropy indexes (AB, AG, and AU) are small and the degree in elastic anisotropy of Al6Mg1Zr1 is relatively weak in the range of εx from 0 to 6%. These results show that applying uniaxial tension strain is an effective and promising strategy to improve the mechanical properties of Al6Mg1Zr1.

Author Contributions

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

Funding

This research was funded by the Natural Science Foundation of Inner Mongolia Autonomous Region (Grant Nos. 2022MS01009 and 2018MS01013), the National Natural Science Foundation of China (Grant Nos. 11972221 and 11562016), the College Science Research Project of Inner Mongolia Autonomous Region (Grant No. NJZY22383), and the Key Research Project of Inner Mongolia University of Technology (Grant No. ZZ202016).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article and can be requested from the corresponding author.

Acknowledgments

The authors acknowledge Mechanics Department of Inner Mongolia University of Technology and the Mathematics Department of Shanghai Maritime University for providing technical support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dolce, D.; Swamy, A.; Hoyt, J.; Choudhury, P. Computing the solid-liquid interfacial free energy and anisotropy of the Al-Mg system using a MEAM potential with atomistic simulations. Comput. Mater. Sci. 2023, 217, 111901. [Google Scholar] [CrossRef]
  2. Xu, W.; Xin, Y.C.; Zhang, B.; Li, X.Y. Stress corrosion cracking resistant nanostructured Al-Mg alloy with low angle grain boundaries. Acta Mater. 2022, 225, 117607. [Google Scholar] [CrossRef]
  3. Koju, R.K.; Mishin, Y. Atomistic study of grain-boundary segregation and grainboundary diffusion in Al-Mg alloys. Acta Mater. 2020, 201, 596–603. [Google Scholar] [CrossRef]
  4. Pariyar, A.; Toth, L.S.; Kailas, S.V.; Peltier, L. Imparting high-temperature grain stability to an Al-Mg alloy. Scr. Mater. 2021, 190, 141–146. [Google Scholar] [CrossRef]
  5. Zhang, H.T.; Guo, C.; Li, S.S.; Li, B.M.; Nagaumi, H. Influence of cold pre-deformation on the microstructure, mechanical properties and corrosion resistance of Zn-bearing 5xxx aluminum alloy. J. Mater. Res. Technol. 2022, 16, 1202–1212. [Google Scholar] [CrossRef]
  6. Bakare, F.; Schieren, L.; Rouxel, B.; Jiang, L.; Langan, T.; Kupke, A.; Weiss, M.; Dorin, T. The impact of L12 dispersoids and strain rate on the Portevin-Le-Chatelier effect and mechanical properties of Al–Mg alloys. Mat. Sci. Eng. A Struct. 2021, 811, 141040. [Google Scholar] [CrossRef]
  7. Yi, G.; Cullen, D.A.; Littrell, K.C.; Golumbfskie, W.; Sundberg, E.; Free, M.L. Characterization of Al-Mg alloy aged at low temperatures. Metall. Mater. Trans. A 2017, 48, 2040–2050. [Google Scholar] [CrossRef]
  8. Zhang, L.; Liu, C.Y.; Zhang, B.; Huang, H.F.; Xie, H.Y.; Cao, K. Mechanical properties of Al–Mg alloys with equiaxed grain structure produced by friction stir processing. Mater. Chem. Phys. 2023, 294, 127010. [Google Scholar] [CrossRef]
  9. Guo, C.; Zhang, H.T.; Li, J.H. Influence of Zn and/or Ag additions on microstructure and properties of Al-Mg based alloys. J. Alloys Compd. 2022, 904, 163998. [Google Scholar] [CrossRef]
  10. Stemper, L.; Tunes, M.A.; Paul, O.; Uggowitzer, P.J.; Pogatscher, S. Age-hardening response of AlMgZn alloys with Cu and Ag additions. Acta Mater. 2020, 195, 541–554. [Google Scholar] [CrossRef]
  11. Zhu, Z.X.; Jiang, X.X.; Wei, G.; Fang, X.G.; Zhong, Z.H.; Song, K.J.; Han, J.; Jiang, Z.Y. Influence of Zn Content on microstructures, mechanical properties and stress corrosion behavior of AA5083 aluminum alloy. Acta Metall. Sin. 2020, 33, 1369–1378. [Google Scholar] [CrossRef]
  12. Chen, X.L.; Marioara, C.D.; Andersen, S.J.; Friis, J.; Lervik, A.; Holmestad, R.; Kobayashi, E. Precipitation processes and structural evolutions of various GPB zones and two types of S phases in a cold- rolled Al-Mg-Cu alloy. Mater. Des. 2021, 199, 109425. [Google Scholar] [CrossRef]
  13. Jiang, L.; Zhang, Z.F.; Bai, Y.L.; Li, S.L.; Mao, W.M. Study on Sc microalloying and strengthening mechanism of Al-Mg alloy. Crystals 2022, 12, 673. [Google Scholar] [CrossRef]
  14. Zakharov, V.V.; Filatov, Y.A.; Fisenko, I.A. Scandium alloying of aluminum alloys. Met. Sci. Heat Treat. 2020, 62, 518–523. [Google Scholar] [CrossRef]
  15. Yan, K.; Chen, Z.W.; Lu, W.J.; Zhao, Y.N.; Le, W.; Naseem, S. Nucleation and growth of Al3Sc precipitates during isothermal aging of Al-0.55 wt% Sc alloy. Mater. Charact. 2021, 179, 111331. [Google Scholar] [CrossRef]
  16. Vo, N.Q.; Dunand, D.C.; Seidman, D.N. Atom probe tomographic study of a friction stir-processed Al-Mg-Sc alloy. Acta Mater. 2012, 60, 7078–7089. [Google Scholar] [CrossRef]
  17. Deng, Y.; Zhu, X.W.; Lai, Y.; Guo, Y.F.; Fu, L.; Xu, G.F.; Huang, J.W. Effects of Zr/(Sc+Zr) microalloying on dynamic recrystallization, dislocation density and hot workability of Al−Mg alloys during hot compression deformation. Trans. Nonferr. Metal. Soc. 2023, 33, 668–682. [Google Scholar] [CrossRef]
  18. Croteau, J.R.; Jung, J.G.; Whalen, S.A.; Darsell, J.; Mello, A.; Holstine, D.; Lay, K.; Hansen, M.; Dunand, D.C.; Vo, N.Q. Ultrafine-grained Al-Mg-Zr alloy processed by shear-assisted extrusion with high thermal stability. Scr. Mater. 2020, 186, 326–330. [Google Scholar] [CrossRef]
  19. Kaiser, M.S.; Rashed, H.M. Precipitation behavior prior to hot roll in the cold-rolled Al-Mg and Al-Mg-Zr alloys. Kov. Mater. 2022, 60, 247–256. [Google Scholar]
  20. Xu, R.; Li, R.D.; Yuan, T.C.; Zhu, H.B.; Wang, M.B.; Li, J.F.; Zhang, W.; Cao, P. Laser powder bed fusion of Al–Mg–Zr alloy: Microstructure, mechanical properties and dynamic precipitation. Mat. Sci. Eng. A Struct. 2022, 859, 144181. [Google Scholar] [CrossRef]
  21. Croteau, J.R.; Griffiths, S.; Rossell, M.D.; Leinenbach, C.; Kenel, C.; Jansen, V.; Seidman, D.N. Microstructure and mechanical properties of Al-Mg-Zr alloys processed by selective laser melting. Acta Mater. 2018, 153, 35–44. [Google Scholar] [CrossRef]
  22. Cho, H.; Lee, B.; Jang, D.; Yoon, J.; Chung, S.; Hong, Y. Recent progress in strain-engineered elastic platforms for stretchable thin-film devices. Mater. Horiz. 2022, 9, 2053–2075. [Google Scholar] [CrossRef] [PubMed]
  23. Chen, Y.M.; Lei, Y.S.; Li, Y.H.; Yu, Y.G.; Cai, J.Z.; Chiu, M.-H.; Rao, R.; Gu, Y.; Wang, C.; Choi, W.; et al. Strain engineering and epitaxial stabilization of halide perovskites. Nature 2020, 577, 209–215. [Google Scholar] [CrossRef] [PubMed]
  24. Dang, C.Q.; Chou, J.-P.; Dai, B.; Chou, C.-T.; Yang, Y.; Fan, R.; Lin, W.T.; Meng, F.; Hu, A.; Zhu, J.; et al. Achieving large uniform tensile elasticity in microfabricated diamond. Science 2021, 371, 76–78. [Google Scholar] [CrossRef] [PubMed]
  25. Steuer, O.; Schwarz, D.; Oehme, M.; Schulze, J.; Mączko, H.; Kudrawiec, R.; Fischer, I.A.; Heller, R.; Hübner, R.; Khan, M.M.; et al. Band-gap and strain engineering in GeSn alloys using post-growth pulsed laser melting. J. Phys. Condens. Mat. 2023, 35, 055302. [Google Scholar] [CrossRef] [PubMed]
  26. Fu, H.; He, W.H.; Mi, Z.S.; Yan, Y.; Zhang, J.L.; Li, J.X. Effects of heterogeneous plastic strain on hydrogen-induced cracking of twinning-induced plasticity steel: A quantitative approach. Corros. Sci. 2022, 209, 110782. [Google Scholar] [CrossRef]
  27. Liang, H.; He, R.Q.; Lin, W.T.; Liu, L.; Xiang, X.J.; Zhang, Z.G.; Guan, S.X.; Peng, F.; Fang, L. Strain-induced strengthening in superconducting β-Mo2C through high pressure and high temperature. J. Eur. Ceram. Soc. 2023, 43, 88–98. [Google Scholar] [CrossRef]
  28. Du, Y.-X.; Zhou, L.-J.; Guo, J.-G. Investigation on micro-mechanism of strain-induced and defect-regulated negative Poisson’s ratio of grapheme. Mater. Chem. Phys. 2022, 288, 126412. [Google Scholar] [CrossRef]
  29. Rasidul Islam, M.; Rayid Hasan Mojumder, M.; Moshwan, R.; Jannatul Islam, A.S.M.; Islam, M.A.; Shizer Rahman, M.; Humaun Kabir, M. Strain-driven optical, electronic, and mechanical properties of inorganic halide perovskite CsGeBr3. ECS J. Solid State Sci. Technol. 2022, 11, 033001. [Google Scholar] [CrossRef]
  30. Tan, Y.; Ma, L.M.; Wang, Y.S.; Zhou, W.; Wang, X.L.; Guo, F. Mechanical and thermodynamic behaviors of AlSi2Sc2 under uniaxial tensile loading: A first-principles study. J. Phys. Chem. Solids 2023, 174, 111160. [Google Scholar] [CrossRef]
  31. Clark, S.J.; Segall, M.D.; Pickard, C.J.; Hasnip, P.J.; Probert, M.I.J.; Refson, K.; Payne, M.C. First Principles Methods Using CASTEP. Z. Krist. Cryst. Mater. 2005, 220, 567–570. [Google Scholar] [CrossRef]
  32. Hohenberg, P.; Kohn, W. Inhomogeneous electron gas. Phys. Rev. 1964, 136, B864–B871. [Google Scholar] [CrossRef]
  33. Kohn, W.; Sham, L.J. Self-consistent equations including exchange and correlation effects. Phys. Rev. 1965, 140, A1133–A1138. [Google Scholar] [CrossRef]
  34. Perdew, J.P.; Burke, K.; Ernzerhof, K.M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [PubMed]
  35. Yang, J.Z.; Yang, D.F.; Wang, Y.Q.; Quan, X.J.; Li, Y.Y. First principles investigation of elastic and thermodynamic properties of CoSbS thermoelectric material. J. Solid State Chem. 2021, 302, 122443. [Google Scholar] [CrossRef]
  36. Mouhat, F.; Coudert, F.-X. Necessary and sufficient elastic stability conditions in various crystal systems. Phys. Rev. B 2014, 90, 224104. [Google Scholar] [CrossRef]
  37. Peng, M.J.; Wang, R.F.; Wu, Y.J.; Yang, A.C.; Duan, Y.H. Elastic anisotropies, thermal conductivities and tensile properties of MAX phases Zr2AlC and Zr2AlN: A first-principles calculation. Vacuum 2022, 196, 110715. [Google Scholar] [CrossRef]
  38. Li, L.H.; Wang, W.L.; Wei, B. First-principle and molecular dynamics calculations for physical properties of Ni-Sn alloy system. Comput. Mater. Sci. 2015, 99, 274–284. [Google Scholar] [CrossRef]
  39. Chen, X.Q.; Niu, H.Y.; Li, D.Z.; Li, Y.Y. Modeling hardness of polycrystalline materials and bulk metallic glasses. Intermetallics 2011, 19, 1275–1281. [Google Scholar] [CrossRef]
  40. Pugh, S.F. XCII. Relations between the elastic moduli and the plastic properties of polycrystalline pure metals. Lond. Edinb. Dublin Philos. Mag. J. Sci. 1954, 45, 823–843. [Google Scholar] [CrossRef]
  41. Zhu, H.Y.; Shi, L.W.; Li, S.Q.; Zhang, S.B.; Xia, W.S. Pressure effects on structural, electronic, elastic and lattice dynamical properties of XSi2 (X = Cr, Mo, W) from first principles. Int. J. Mod. Phys. B 2018, 32, 1850120. [Google Scholar] [CrossRef]
  42. Yang, A.C.; Bao, L.K.; Peng, M.J.; Duan, Y.H. Explorations of elastic anisotropies and thermal properties of the hexagonal TMSi2 (TM = Cr, Mo, W) silicides from first-principles calculations. Mater. Today Commun. 2021, 27, 102474. [Google Scholar] [CrossRef]
  43. Zhu, H.Y.; Shi, L.W.; Li, S.Q.; Zhang, S.B.; Xia, W.S. Effects of biaxial strains on electronic and elastic properties of hexagonal XSi2 (X = Cr, Mo, W) from first-principles. Solid State Commun. 2018, 270, 99–106. [Google Scholar] [CrossRef]
Figure 1. Structural model of Al6Mg1Zr1 supercell.
Figure 1. Structural model of Al6Mg1Zr1 supercell.
Metals 13 01569 g001
Figure 2. Stress–strain relations of Al6Mg1Zr1 at different uniaxial tension strains in x direction (εx).
Figure 2. Stress–strain relations of Al6Mg1Zr1 at different uniaxial tension strains in x direction (εx).
Metals 13 01569 g002
Figure 3. Elastic moduli (B, G, and E) and Poisson ratio μ of Al6Mg1Zr1 at different uniaxial tension strains (εx): (a) Bulk modulus B vs. strain εx; (b) Shear modulus G vs. strain εx; (c) Young‘s modulus E vs. strain εx; (d) Poisson ratio μ vs. strain εx.
Figure 3. Elastic moduli (B, G, and E) and Poisson ratio μ of Al6Mg1Zr1 at different uniaxial tension strains (εx): (a) Bulk modulus B vs. strain εx; (b) Shear modulus G vs. strain εx; (c) Young‘s modulus E vs. strain εx; (d) Poisson ratio μ vs. strain εx.
Metals 13 01569 g003
Figure 4. Vickers hardness Hv and the modulus ratio B/G of Al6Mg1Zr1 at different uniaxial tension strain (εx): (a) Vickers hardness Hv vs. strain εx; (b) the ratio D (D = B/G) vs. strain εx.
Figure 4. Vickers hardness Hv and the modulus ratio B/G of Al6Mg1Zr1 at different uniaxial tension strain (εx): (a) Vickers hardness Hv vs. strain εx; (b) the ratio D (D = B/G) vs. strain εx.
Metals 13 01569 g004
Figure 5. Elastic anisotropy indexes of Al6Mg1Zr1 (AB, AG, and AU) at different uniaxial tension strains (εx): (a) AB vs. strain εx; (b) AG vs. strain εx; (c) AU vs. strain εx.
Figure 5. Elastic anisotropy indexes of Al6Mg1Zr1 (AB, AG, and AU) at different uniaxial tension strains (εx): (a) AB vs. strain εx; (b) AG vs. strain εx; (c) AU vs. strain εx.
Metals 13 01569 g005
Table 1. Calculated values of the elastic stiffness constants Cij (in GPa) of Al6Mg1Zr1 at different uniaxial tension strains in x direction (εx).
Table 1. Calculated values of the elastic stiffness constants Cij (in GPa) of Al6Mg1Zr1 at different uniaxial tension strains in x direction (εx).
εx (%)C11C12C13C22C23C33C44C55C66
0 157.17 41.58 41.56 170.16 41.57 170.08 40.52 53.70 53.71
1%146.98 36.34 36.37 163.72 42.30 163.79 42.19 49.69 49.90
2%136.71 35.30 35.32 156.80 45.54 156.82 44.84 47.06 47.06
3%124.44 35.59 35.57 146.39 48.13 146.36 48.54 43.25 43.25
4%111.37 37.46 37.46 133.26 50.68 133.26 51.02 37.65 37.65
5%94.62 39.78 39.78 118.19 55.44 118.19 50.42 29.91 29.91
6%65.16 44.63 44.63 99.25 64.20 99.26 46.10 19.98 19.98
7%17.81 52.18 52.20 75.93 75.41 75.95 40.10 8.17 8.17
Table 2. Calculated values of stresses in principal axis directions (σx, σy, and σz) of Al6Mg1Zr1 at different uniaxial tension strains in x direction (εx).
Table 2. Calculated values of stresses in principal axis directions (σx, σy, and σz) of Al6Mg1Zr1 at different uniaxial tension strains in x direction (εx).
εx (%)σx (GPa)σy (GPa)σz (GPa)
0000
1%1.570.420.42
2%3.040.780.78
3%4.411.131.13
4%5.651.491.49
5%6.771.87 1.86
6%7.712.262.26
Table 3. Calculated values of the elastic moduli (B, G, and E) and Poisson ratios μ of polycrystalline Al6Mg1Zr1 at different uniaxial tension strains in x direction (εx).
Table 3. Calculated values of the elastic moduli (B, G, and E) and Poisson ratios μ of polycrystalline Al6Mg1Zr1 at different uniaxial tension strains in x direction (εx).
εx (%)B (GPa)G (GPa)E (GPa)μ
082.93 53.77132.63 0.23
1%78.12 51.81 127.28 0.23
2%75.55 49.82 122.53 0.230
3%72.43 46.72 115.36 0.24
4%69.37 41.76 104.34 0.25
5%65.94 34.36 87.81 0.28
6%60.63 22.88 60.96 0.33
Table 4. Calculated values of the Vickers hardness HV and the modulus ratio B/G of Al6Mg1Zr1 at different uniaxial tension strains in x direction (εx).
Table 4. Calculated values of the Vickers hardness HV and the modulus ratio B/G of Al6Mg1Zr1 at different uniaxial tension strains in x direction (εx).
εx (%)HV (GPa)B/G
0 11.97 1.54
1%12.03 1.51
2%11.67 1.52
3%10.95 1.55
4%9.45 1.66
5%7.11 1.92
6%3.83 2.65
Table 5. Calculated values of elastic anisotropy indexes (AB, AG, and AU) of Al6Mg1Zr1 at different uniaxial tension strains in x direction (εx).
Table 5. Calculated values of elastic anisotropy indexes (AB, AG, and AU) of Al6Mg1Zr1 at different uniaxial tension strains in x direction (εx).
εx (%)ABAGAU
0 0.06%1.26%0.13
1%0.21% 0.92%0.10
2%0.40%0.50%0.06
3%0.60%0.32% 0.04
4%0.77%0.77% 0.09
5%1.27%2.20%0.25
6%4.58%7.37%0.89
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

Zhang, L.; Li, J.; Zhang, J.; Liu, Y.; Lin, L. First-Principle Investigation into Mechanical Properties of Al6Mg1Zr1 under Uniaxial Tension Strain on the Basis of Density Functional Theory. Metals 2023, 13, 1569. https://doi.org/10.3390/met13091569

AMA Style

Zhang L, Li J, Zhang J, Liu Y, Lin L. First-Principle Investigation into Mechanical Properties of Al6Mg1Zr1 under Uniaxial Tension Strain on the Basis of Density Functional Theory. Metals. 2023; 13(9):1569. https://doi.org/10.3390/met13091569

Chicago/Turabian Style

Zhang, Lihua, Jijun Li, Jing Zhang, Yanjie Liu, and Lin Lin. 2023. "First-Principle Investigation into Mechanical Properties of Al6Mg1Zr1 under Uniaxial Tension Strain on the Basis of Density Functional Theory" Metals 13, no. 9: 1569. https://doi.org/10.3390/met13091569

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