Acta Metallurgica Sinica(English Letters), 2017, 33(5): 475-480
doi: 10.1016/j.jmst.2016.10.003
Effect of Icosahedral Phase on Crystallographic Texture and Mechanical Anisotropy of Mg-4%Li Based Alloys
C.Q. Li1,2, D.K. Xu2,*,, S. Yu2,3, L.Y. Sheng4, E.H. Han2

Abstract:

Through investigating and comparing the microstructure and mechanical properties of the as-extruded Mg alloys Mg-4%Li and Mg-4%Li-6%Zn-1.2%Y (in wt%), it demonstrates that although the formation of I-phase (Mg3Zn6Y, icosahedral structure) could weaken the crystallographic texture and improve the mechanical strength, the mechanical anisotropy in terms of strength remains in Mg-4%Li-6%Zn-1.2%Y alloy. Failure analysis indicates that for the Mg-4%Li alloy, the fracture surfaces of the tensile samples tested along transverse direction (TD) contain a large number of plastic dimples, whereas the fracture surface exhibits quasi-cleavage characteristic when tensile samples were tested along extrusion direction (ED). For the Mg-4%Li-6%Zn-1.2%Y alloy, typical ductile fracture surfaces can be observed in both “TD” and “ED” samples. Moreover, due to the zonal distribution of broken I-phase particles, the fracture surface of “TD” samples is characterized by the typical “woody fracture”.

Key words: Mg-Li alloy; Texture; Mechanical anisotropy; Fracture;
1. Introduction

It is well known that crystallographic texture plays a critical role in mechanical properties and formability of wrought magnesium (Mg) alloys[1,2,3]. Generally, addition of lithium (Li) in Mg matrix is an effective way to weaken their texture. With the increase in Li content, the c/a axial ratio decreases from 1.624 for pure Mg to 1.607 for Mg-17 at.% Li [4,5,6]. Due to the decreased c/a ratio, the rotation of basal poles to the transverse direction can be promoted [4,7]. Compared with commercial Mg alloy AZ31, Mg-4%Li alloy could exhibit weaker basal and prismatic textures through an identical cold-rolling process[5].

Although Mg-Li alloys could have a weakened crystallographic texture, their mechanical anisotropy could retain[5,7,8]. Compared with traditional Mg alloys, their mechanical strength is obviously lower[8,9]. Recently, Xu et al. reported that through adding Zn and Y and controlling Zn/Y atomic ratio of 6, I-phase (Mg3Zn6Y, icosahedral quasicrystal structure) could be formed in Mg-Li alloys[8,10,11]. It demonstrated that the formation of I-phase in duplex structured Mg-Li alloys could not only improve their mechanical strength, but also weaken their crystallographic texture and mechanical anisotropy[11]. Generally, for metallic materials, the particle stimulated nucleation (PSN) of recrystallization could be beneficial for weakening the texture due to the recrystallization nuclei at particles having a random crystallographic orientation[12,13,14,15]. Similarly, the basal texture of wrought Mg alloys can also be weakened by the PSN[16,17,18,19]. In addition, the main reason for the formation of weaker basal texture in Mg-6%Li-6%Zn-1.2%Y alloy is that the broken I-phase particles can induce the formation of stress concentration zones with a high density of dislocations and accelerate the recrystallization process of the surrounding α-Mg matrix during the severe plastic deformation[11]. Karami et al. reported that the formed β-Li phases could also be beneficial for weakening the basal texture of α-Mg phases in duplex structured Mg-Li alloys, which was ascribed to the fact that the incoherent interface and different orientation between α-Mg and β-Li phases could restrict the preferential reorientation of the neighboring recrystallized α-Mg grains during deformation[3]. Therefore, for I-phase containing Mg-Li alloys, it is challenging to identify the individual contribution of PSN and β-Li phases to the weakening in texture and mechanical anisotropy. In this work, through investigating and comparing the microstructure and mechanical properties of single structured Mg-4%Li and Mg-4%Li-6%Zn-1.2%Y alloys, the effect of I-phase-formation derived PSN on the crystallographic texture and mechanical anisotropy was explored.

2. Experimental Procedures

The materials used in this study were as-extruded Mg-4%Li and Mg-4%Li-6%Zn-1.2%Y (in wt%) alloys prepared at the Institute of Metal Research, Chinese Academy of Sciences. Through inductively coupled plasma atomic emission spectrum (ICP-AES) apparatus, the chemical composition of the alloys was determined (Table 1). The extrusion ratio was 6:1.


Table 1. Chemical composition of the Mg-4%Li and Mg-4%Li-6%Zn-1.2%Y alloys
Alloys Li Zn Y Mg
Mg-4%Li 4.2 Bal
Mg-4%Li-6%Zn-1.2%Y 4.3 5.8 1.2 Bal

Table 1. Chemical composition of the Mg-4%Li and Mg-4%Li-6%Zn-1.2%Y alloys

Sample pieces cut from the as-extruded plates were ground with SiC paper progressively up to 5000 grit finish, and finely polished to a 1 µm finish with ethanol. Microstructure was observed by optical microscopy (OM), scanning electron microscopy (SEM; XL30-FEG-ESEM), electron back-scattering diffraction (EBSD) and transmission electron microscopy (TEM; JEOL 2100) equipped with energy dispersive X-ray spectroscopy (EDS). Thin foil specimens for TEM observation were prepared by mechanical thinning, followed by argon ion milling. To reveal the difference in grain structure between the alloys, polished samples were etched with 4% nitric acid + 96% ethanol. Average grain size was determined by using the mean linear intercept method. Distribution intensity of the (0002) pole figures for the two alloys was measured by EBSD analysis. Tensile samples with a gauge length of 25 mm, width of 5 mm and thickness of 3 mm were machined from the as-extruded plates. Moreover, samples with axial direction parallel to the transverse and extrusion directions of the plates were defined as “TD” and “ED” samples, respectively, as shown in Fig. 1. Tensile tests were conducted on a MTS (858.01 M) testing machine at a constant strain rate of 1 × 10-3 s-1 at room temperature. After tensile testing, the fracture surfaces and side surfaces near the fracture areas were observed by SEM.

Fig. 1. Schematic diagram of tensile samples.

3. Results
3.1. Microstructural characterization

Since Mg-4%Li alloy is only composed of α-Mg matrix[5], only the microstructure of Mg-4%Li-6%Zn-1.2%Y alloy is provided, as shown in Fig. 2. It demonstrates that a great number of coarse particles distribute along the extrusion direction and a high density of fine particles scatter in the matrix, as shown in Fig. 2(a) and (b). To identify these particles, their bright field images and corresponding selected area diffraction patterns (SADPs) were obtained, as shown in Fig. 2(c-f). It reveals that the coarse particles exhibit a typical 5-fold symmetry, as shown in Fig. 2(d). This result further confirms the existence of I-phase pockets in Mg-4%Li-6%Zn-1.2%Y alloy. Moreover, it is recognized that for Mg alloys containing 4-9 wt% Zn and aged isothermally at 120-260 °C, the strengthening precipitates are meta-stable β1′ and β2′, respectively[20,21,22,23,24]. Previous work reported that the β1′ and β2′ precipitates are either MgZn2 having a hexagonal structure (a = 0.520 nm, c = 0.857 nm) [25] or Mg4Zn7 phase having a base-centered monoclinic structure (a = 2.596 nm, b = 1.428 nm, c = 0.524 nm, γ = 102.5°) [20]. Additionally, the long axis of rod-like β1′ is parallel to the [0002]α-Mg, whereas the β2′ precipitates have an orientation relationship of (0002)α-Mg // (0002)β2′ and [10-10]α-Mg // [11-20]β2′ with the α-Mg matrix[21]. Following this, the fine precipitates in the matrix can be determined as β1′ and β2′ precipitates, as shown in Fig. 2(e) and (f). Thus, the alloy is mainly composed of α-Mg, I-phase and meta-stable MgZn precipitates.

Fig. 2. Microstructures of as-extruded Mg-4%Li-6%Zn-1.2%Y alloy: (a) SEM image, (b) high magnification observation to the squared area of image (a), (c) TEM bright field image and (d) 5-fold SADP of I-phase, (e) TEM bright field image of β1′, β2′, (f) SADP of α-Mg and β2′.

Fig. 3 shows the grain structure of the two alloys. It can be seen that for Mg-4%Li alloy, the grain size is heterogeneous and varies from 5 µm to 200 µm, which can be attributed to an incomplete recrystallization. For Mg-4%Li-6%Zn-1.2%Y alloy, the recrystallization could be completed, resulting in a relatively uniform grain structure with an average grain size of 25 µm. To further confirm this, EBSD analysis to the grain structure of the two alloys was performed, as shown in Fig. 4. It can be seen that the obtained grain structure of the two alloys (Fig. 4(a) and (b)) is consistent with that of the optical images (Fig. 3(a) and (b)). The measured recrystallized fraction in the two alloys is shown in Fig. 4(c-f). It reveals that for the grain structure of Mg-4%Li alloy, the fraction of deformed grains is 90% and the recrystallized fraction is less than 5% (Fig. 4(e)). After the formation of I-phase, the fraction of recrystallized grains in Mg-4%Li-6%Zn-1.2%Y alloy approaches 33% and the deformed fraction decreases to 31% (Fig. 4(f)). Fig. 5 shows the measured (0002) pole figures of two alloys. It demonstrates that the crystallographic texture in the two alloys is similar, i.e. the c-axis of most grains is approximately perpendicular to the extrusion direction (ED) and inclines 45° towards the transverse direction (TD). However, the maximum intensity of texture in Mg-4%Li-6%Zn-1.2%Y alloy is obviously weaker than that of Mg-4%Li alloy, which can be ascribed to the PSN effect of I-phase particles during hot extrusion[26,27].

Fig. 3. Optical observations on the etched (a) Mg-4%Li and (b) Mg-4%Li-6%Zn-1.2%Y samples.

Fig. 4. EBSD results of the as-extruded alloys: (a) and (b) orientation maps, (c) and (d) recrystallized maps, (e) and (f) distribution of recrystallized fraction for Mg-4%Li and Mg-4%Li-6%Zn-1.2%Y samples, respectively.

Fig. 5. (0002) pole figures of the as-extruded alloys: (a) Mg-4%Li and (b) Mg-4%Li-6%Zn-1.2%Y.

3.2. Mechanical properties

Typical engineering stress-strain curves of the two alloys are shown in Fig. 6. To describe and compare the measured data, the mechanical properties of 0.2% proof yield strength (YS), ultimate tensile strength (UTS) and elongation ratio to failure (EL) of the two alloys are listed in Table 2. It indicates that for the Mg-4%Li alloy, the tensile strength along the “TD” direction is lower than that along the “ED” direction (Fig. 6(a)). YS and UTS of the “TD” sample are 80 MPa and 141 MPa, whereas the corresponding values of the “ED” sample are 105 MPa and 163 MPa. However, EL of the “TD” sample is 1.5 times higher than that of the “ED” sample. For the Mg-4%Li-6%Zn-1.2%Y alloy, the strength anisotropy is still apparent, but EL values of the two oriented samples are basically the same (Fig. 6(b)). Moreover, YS and UTS of Mg-4%Li-6%Zn-1.2%Y alloy are 23-36 MPa and 77-85 MPa higher than those of Mg-4%Li alloy. Therefore, it demonstrates that the formation of I-phase is beneficial to mechanical strength, which is consistent with previous results[8,10,11].

Fig. 6. Engineering stress-strain curves of as-extruded alloys: (a) Mg-4%Li and (b) Mg-4%Li-6%Zn-1.2%Y.


Table 2. Tensile properties of as-extruded Mg-4%Li and Mg-4%Li-6%Zn-1.2%Y alloys
Alloys Orientation YS (MPa) UTS (MPa) EL (%)
Mg-4%Li TD sample 80 ± 5 141 ± 3 25.7 ± 2
ED sample 105 ± 4 163 ± 4 16.5 ± 3
Mg-4%Li-6%Zn-1.2%Y TD sample 103 ± 3 226 ± 5 22.8 ± 2
ED sample 141 ± 5 240 ± 5 21.6 ± 3

Table 2. Tensile properties of as-extruded Mg-4%Li and Mg-4%Li-6%Zn-1.2%Y alloys

3.3. Failure analysis and fractography

To disclose the deformation mechanisms that occurred in differently oriented samples, the microstructure of side surfaces with a distance of about 1 mm from the fracture site was observed, as shown in Fig. 7. It can be seen that for the Mg-4%Li alloy, the main deformation mechanism of the “TD” sample is dominated by the dislocation slips (Fig. 7(a)), whereas the twinning can be widely activated in the “ED” sample (Fig. 7(b)). For the Mg-4%Li-6%Zn-1.2%Y alloy, the dislocation slips can be observed on the surfaces of both “TD” and “ED” samples (Fig. 7(c) and (d)). However, numerous micro cracks exist in the I-phase particles when tensile was tested along the “TD” direction (Fig. 7(c)).

Fig. 7. Observations on the side surfaces of differently oriented samples: (a) TD and (b) ED samples cut from Mg-4%Li alloy, (c) TD and (d) ED samples cut from Mg-4%Li-6%Zn-1.2%Y alloy.

To further confirm the activation of twinning in different samples, the etched microstructure of the side surfaces was examined (Fig. 8). It clearly demonstrates that for both alloys, the twinning in the “ED” sample is much more active than that in the “TD” sample. Fig. 9 shows the fracture surfaces of differently oriented samples. It can be seen that for Mg-4%Li alloy, a lot of plastic dimples exist on the fracture surface of “TD” sample (Fig. 9(a)), whereas the fracture surface of “ED” sample contains many cleavage planes (Fig. 9(b)). For the Mg-4%Li-6%Zn-1.2%Y alloy, the fracture surfaces of the two oriented samples are similar and mainly composed of plastic dimples, as shown in Fig. 9(c) and (d). Meanwhile, a large number of secondary phase particles exist at the bottom of dimples (Fig. 9(e) and (f)). Due to the zonal distribution of I-phase particles along the extrusion direction, the “woody fracture” characteristic occurs on the fracture surfaces for the “TD” samples.

Fig. 8. Observations on the etched side surfaces of TD (a) and ED (b) samples cut from Mg-4%Li alloy, TD (c) and ED (d) samples cut from Mg-4%Li-6%Zn-1.2%Y alloy.

Fig. 9. Observations on the fracture surfaces of TD (a) and ED (b) samples of Mg-4%Li alloy; TD (c) and ED (d) samples of Mg-4%Li-6%Zn-1.2%Y alloy; images (e) and (f) are backscatter secondary images in-situ of images (c) and (d).

4. Discussion

Generally, icosahedral phase (I-phase) exhibits many excellent properties such as high hardness, thermal stability, high corrosion resistance, low coefficient of friction, and low interfacial energy[28]. Since I-phase could have semi-coherent interfaces with the α-Mg matrix by introducing steps and ledges periodically along the interface[29], the I-phase/α-Mg interfaces could effectively retard the basal slip of dislocations[30]. Moreover, Bae et al. reported that the I-phase/α-Mg interfaces could exhibit strong thermal stability and effectively resist the microstructural coarsening at elevated temperatures[29,31]. Furthermore, the in-situ formation of I-phase in Mg-Li alloys can effectively improve their mechanical strength and plasticity[10,11,28]. During hot deformation process, I-phase particles could act as nucleation sites and stimulate the dynamic recrystallization[11]. Generally, large-sized particles (>1 µm) could promote nucleation, which was called “particle stimulated nucleation (PSN)”, but the small-sized or closely spaced particles (with the inter space being less than 0.1 µm) could retard the nucleation[32]. Similarly, Robson et al. reported that the fine dispersed nanoscale particles had little effect on the dynamic recrystallization fraction in magnesium-manganese alloys during the channel die deformation process[33]. Therefore, for the investigated Mg-4%Li-6%Zn-1.2%Y alloy, the effect of the existed nanoscale β1′ and β2′ precipitates on the dynamic recrystallization of the α-Mg matrix during hot extrusion could be neglected. Then, the weaker crystallographic texture in Mg-4%Li-6%Zn-1.2%Y alloy can be mainly attributed to the promoted recrystallization by the existed I-phase particles.

For a strong textured Mg alloy, Schmidt factor (SF) can be expressed as [34]: SF = cosΦcosλ, where Φ and λ (angles Φ and λ are complementary (Φ + λ = 90°)) are the angles of tensile direction with respect to the normal direction of slip plane and optimal slip direction, respectively. For the Mg-4%Li alloy, the angles between c-axis and tensile direction of “TD” samples are mainly concentrated about 45°, but the same value of “ED” samples presents an angle about 80°-100° ( Fig. 5(a)). In addition, the critical resolved shear stress (CRSS) of deformation models with the following relationship [35,36]: CRSSbasal < CRSStwinning < CRSSprismatic < CRSSpyramidal. As a result, the SF of basal plane for “TD” sample is about 0.5, but the SF of the “ED” sample is lower than 0.17. Thus, the basal slip occurs easily for the “TD” sample and it ensures a good plasticity, whereas the basal slip can be hardly activated due to the low SF of the “ED” samples. In general, {10-12} twins were profuse in magnesium alloys and could be activated when the tension was along the c-axis or compression was perpendicular to the c-axis of the grain [37,38]. Moreover, even though c-axis of grains is parallel to the tensile direction, the activation of some {10-12} extension twins can still occur [39,40]. Therefore, for the “ED” samples, the {10-12} extension twinning forms easily to make up for the lack of slip systems at room temperature (Figs. 7(b) and 8(b)).

Serrated flow phenomena of differently oriented samples are observed in two alloys (Fig. 6). Previous work demonstrated that the occurrence of serrated flow in Mg alloys was mainly ascribed to the dynamic strain aging (DSA), i.e. the interaction between solute atoms and moveable dislocations[41,42]. Thus, the effect of I-phase particles on the serration phenomenon of Mg-4%Li-6%Zn-1.2%Y alloy can be neglected. Moreover, Wang et al. found that the solute atoms could diffuse and even pin the dislocations when the diffusion velocity of solute atoms was equal to the dislocation mobility, resulting in the increase in tensile stress[42]. Thus, when the applied tensile stress exceeds a certain level, the pinned dislocations in two alloys can escape from the solute atoms and then cause a sudden drop of the stress. With the further increase in the applied tensile stress, the repeated process of pinning and unpinning of dislocations at solute atoms can occur, resulting in the formation of serrated flowing curves (Fig. 6). However, compared with the Mg-4%Li alloy, the amplitude of serrated flowing stress of the Mg-4%Li-6%Zn-1.2%Y alloy is much bigger, which is probably ascribed to the added elements of Zn and Y. Corby et al. reported that Zn atoms could act as a catalyzer and promote the formation of forest dislocations for the pipe diffusion of Al atoms, leading to the increased serration phenomenon of AZ91 Mg alloy[43]. Moreover, Gao et al. found that with the yttrium content increasing, the serrations of binary Mg-Y alloys became more and more pronounced[44]. Therefore, the combined effect of added Zn and Y contributes to more pronounced serrations of Mg-4%Li-6%Zn-1.2%Y alloy.

In terms of texture, Mg-4%Li-6%Zn-1.2%Y alloy presents similar characteristics to that of Mg-4%Li alloy. Moreover, the coarse I-phase particles preferentially distribute along the “ED” direction (Fig. 2(a)), leading to a similar “fiber-strengthening” effect on the “ED” sample. However, when tensile was tested along the “TD” direction, the zonal distribution of course I-phase particles can easily induce the stress concentration and then cause the cracking of them (Fig. 7(c)). Therefore, the strength anisotropy can also exist in Mg-4%Li-6%Zn-1.2%Y alloy. However, the cracking of I-phase particles can accelerate the fracture process even the “TD” sample having favorable orientation for the activation of basal and non-basal slips. Consequently, the anisotropy in elongation between two oriented samples can hardly be observed.

5. Conclusion

Through investigating the crystallographic texture and mechanical anisotropy of two Mg-4%Li based alloys, it demonstrates that the mechanical strength of as-extruded Mg-4%Li-6%Zn-1.2%Y alloy is much higher than that of Mg-4%Li alloy. Although the formation of I-phase can weaken the crystallographic texture, the strength anisotropy remains in Mg-4%Li-6%Zn-1.2%Y alloy. Since the cracking of I-phase particles in “TD” samples occurs and then accelerates the fracture process, the anisotropy in elongation for Mg-4%Li-6%Zn-1.2%Y alloy can be eliminated even the “TD” samples having a favorable orientation for the activation of basal and non-basal slips.

Acknowledgments:This work was supported by the National Natural Science Foundation of China projects under Nos. 51271183, 51171192 and 51301172, the National Basic Research Program of China (973 Program) project under Grant No. 2013CB632205, the National Key Research and Development Program of China project under Grant No. 2016YFB0301105. Shenzhen Technology Innovation Plan (CXZZ20140419114548507 and CXZZ20140731091722497), Shenzhen Basic Research Project (JCYJ20150529162228734) and the Innovation Fund of Institute of Metal Research (IMR), Chinese Academy of Sciences (CAS).

The authors have declared that no competing interests exist.

References

[1] N. Tahreen, D.F. Zhang, F.S. Pan, X.Q. Jiang, D.Y. Li, D.L. Chen, J. Mater.Sci.Technol.31(2015) 1161-1170.
[Cited within:1]
[2] Y.N. Lin, H.Y.Wu, G.Z. Zhou, C.H. Chiu, S. Lee, Mater. Des. 29(2008) 2061-2065.
[Cited within:1]
[3] M. Karami, R. Mahmudi, Metall. Mater. Trans. A 44(2013) 3934-3946.
[Cited within:2]
[4] S.R. Agnew, M.H. Yoo, C.N. Tome, Acta Mater. 49(2001) 4277-4289.
[Cited within:2]
[5] T. Al-Samman, Acta Mater. 57(2009) 2229-2242.
[Cited within:4]
[6] H.T. Son, Y.H. Kim, D.W. Kim, J.H. Kim, H.S. Yu, J. Alloy.Compd. 564(2013)130-137.
[Cited within:1]
[7] R.H. Li, F.S. Pan, B. Jiang, H.W. Dong, Q.S. Yang, Mater. Sci. Eng. A 562 (2013)33-38.
[Cited within:2]
[8] D.K. Xu, L. Liu, Y.B. Xu, E.H. Han, Scr. Mater. 57(2007) 285-288.
[Cited within:4]
[9] J.H. Zhang, L. Zhang, Z. Leng, S.J. Liu, R.Z.Wu, M.L. Zhang, Scr. Mater. 68(2013)675-678.
[Cited within:1]
[10] D.K. Xu, T.T. Zu, M. Yin, Y.B. Xu, E.H. Han, J. Alloy. Compd. 582(2014) 161-166.
[Cited within:3]
[11] D.K. Xu, C.Q. Li, B.J.Wang, E.H. Han, Mater. Des. 88(2015) 88-97.
[Cited within:6]
[12] M. Ferry, F.J. Humphreys, Acta Mater. 44(1996) 3089-3103.
[Cited within:1]
[13] M.G. Ardakani, F.J. Humphreys, Acta Metall.Mater. 42(1994) 763-780.
[Cited within:1]
[14] O. Engler, J. Hirsch, K. Lucke, Acta Metall.Mater. 43(1995) 121-138.
[Cited within:1]
[15] O. Engler, P. Yang, X.W. Kong, Acta Mater. 44(1996) 3349-3369.
[Cited within:1]
[16] S. Kleiner, P.J. Uggowitzer, Mater. Sci. Eng. A 379 (2004) 258-263.
[Cited within:1]
[17] E. Ball, P. Prangnell, Scr. Metall. Mater. 31(1994) 111-116.
[Cited within:1]
[18] J. Bohlen, S.B. Yi, J. Swiostek, D. Letzig, H.G. Brockmeier, K.U. Kainer, Scr. Mater.53(2005) 259-264.
[Cited within:1]
[19] J.R. Dong, D.F. Zhang, J. Sun, Q.W. Dai, F.S. Pan, J. Mater. Sci. Technol. 31(2015)
[Cited within:1]
935-940.
[20] X. Gao, J.F. Nie, Scr. Mater. 56(2007) 645-648.
[Cited within:2]
[21] K. Oh-ishi, K. Hono, K.S. Shin, Mater. Sci. Eng. A 496 (2008) 425-433.
[Cited within:2]
[22] H. Liu, F. Xue, J. Bai, J. Zhou, Y.S. Sun, J. Mater. Sci.Technol. 30(2014) 128-133.
[Cited within:1]
[23] J.M. Rosalie, B.R. Pauw, Acta Mater. 66(2014) 150-162.
[Cited within:1]
[24] C.L. Mendis, K. Oh-ishi, T. Ohkubo, K. Hono, Mater. Sci. Eng. A 535(2012)1122-128.
[Cited within:1]
[25] Y. Komura, K. Tokunaga, Acta Crystallogr. 36(1980) 1548-1554.
[Cited within:1]
[26] H. Chang, X.J.Wang, X.S. Hu, Y.Q.Wang, K.B. Nie, K.Wu, Rare Metal Mater. Eng.A 43 (2014) 1821-1825.
[Cited within:1]
[27] Y.H. Kim, J.H. Kim, H.S. Yu, J.W. Choi, H.T. Son, J. Alloy. Compd. 583(2014) 15-20.
[Cited within:1]
[28] D.K. Xu, E.H. Han, Prog. Nat. Sci. Mater. Int. 22(2012) 364-385.
[Cited within:2]
[29] D.H. Bae, S.H. Kim, D.H. Kim,W.T. Kim, Acta Mater. 50(2002) 2343-2356.
[Cited within:2]
[30] D.K. Xu,W.N. Tang, L. Liu, Y.B. Xu, E.H. Han, J. Alloy. Compd. 432(2007) 129-134.
[Cited within:1]
[31] D.H. Bae, Y. Kim, I.J. Kim, Mater. Lett. 60(2006) 2190-2193.
[Cited within:1]
[32] R.D. Doherty, D.A. Hughes, F.J. Humphreys, J.J. Jonas, D. Juul Jensen, M.E. Kassner,W.E. King, T.R. McNelley, H.J. McQueen, A.D. Rollett, Mater. Sci. Eng. A 238 (1997)219-274.
[Cited within:1]
[33] J.D. Robson, D.T. Henry, B. Davis, Mater. Sci. Eng. A 528(2011) 4239-4247.
[Cited within:1]
[34] W.J. Ding, L. Jin,W.X.Wu, J. Dong, Chin. J. Nonferr.Met. 21(2011) 2371-2381.
[Cited within:1]
[35] T.Z. Wang, T.L. Zhu, R.Z. Wu, W. Miao, J.H. Zhang, M.L. Zhang, Mater. Des. 85(2015) 190-196.
[Cited within:1]
[36] M.R. Barnett, Metall. Mater. Trans. A 34 (2003) 1799-1806.
[Cited within:1]
[37] D.K. Xu, E.H. Han, Scr. Mater. 69(2013) 702-705.
[Cited within:1]
[38] H.E. Kadiri, C.D. Barrett, J.Wang, C.N. Tome, Acta Mater. 85(2015) 354-361.
[Cited within:1]
[39] X. Wang, L. Jiang, A. Luo, J. Song, Z. Liu, F. Yin, Q. Han, S. Yue, J.J. Jonas, J. Alloy.Compd. 594(2014) 44-47.
[Cited within:1]
[40] S.D.Wang, D.K. Xu, B.J.Wang, E.H. Han, C. Dong, Sci. Rep. 6(2016) 23955.
[Cited within:1]
[41] C.Wang, Y.B. Xu, E.H. Han, Mater. Lett. 60(2006) 2941-2944.
[Cited within:1]
[42] W.H.Wang, D.Wu, S.S.A. Shah, R.S. Chen, C.S. Lou, Mater. Sci. Eng. A 649 (2016)214-221.
[Cited within:2]
[43] C. Corby, C.H. Cáceres, P. Luká, Mater. Sci. Eng.A 387-389(2004) 22-24.
[Cited within:1]
[44] L. Gao, R.S. Chen, E.H. Han, J. Alloy. Compd. 472(2009) 234-240.
[Cited within:1]
Downloaded    
RichHTML Viewed    
Abstract Viewed    

Share
Export

Related articles:
Key words
Mg-Li alloy
Texture
Mechanical anisotropy
Fracture

Authors
C.Q. Li
D.K. Xu
S. Yu
L.Y. Sheng
E.H. Han