logo
热轧过程促进非基面滑移开动弱化AZ31镁合金板材平面各向异性研究

热轧过程促进非基面滑移开动弱化AZ31镁合金板材平面各向异性研究

朝阳
利飞
亮亮
秋燕
大彪
新卫
留伟
红霞
KWANG Seon-Shin
3100

为弱化镁合金基面织构和平面各向异性,本文采用高温(350 ℃和450 ℃)单道次大应变(45%)预轧制工艺促进非基面滑移激活开动,随后在300 ℃进行5道次轧制,得到厚度为1 mm镁合金板材。高温预轧后,非基面滑移被促进激活开动,形成分散的双峰织构,从而弱化了预轧镁合金基面织构强度。此外,晶内取向差轴(IGMA,分析滑移激活的一种方法)分布情况表明,锥面滑移在变形过程中具有很强的激活特性。终轧镁板退火后,由于静态再结晶,织构分布发生变化,织构强度明显降低。特别是预轧镁合金样品r值和平面各向异性明显低于原始未预轧样品。

镁合金微观结构各向异性织构演变

J.Cent.South Univ.(2025) 32: 706-726

1 Introduction

Mg alloys are known as the lightest structural materials, and are widely used in the field of 3C electron, automobile and aerospace due to their low density and high specific strength [1-3]. Although wrought Mg alloys have many attractive properties,their applications are still limited due to their strong basal texture [4, 5]. It is well known that the critical resolved shear stress (CRSS) of basal slip is far lower than that of non-basal slip systems at room temperature [6, 7]. Thus, limited by the special hexagonal close-packed crystal structure, only two independent slip systems are activated at room temperature for conventional rolling, which results in an extremely strong basal texture [8, 9]. The strong basal texture gives rise to high anisotropy in the plate and increases the difficulty of strain accommodation in the thickness direction at room temperature [10, 11].

To reduce the anisotropy of Mg alloys, some studies have focused on different processes to weaken the texture. TIAN et al [12] obtained sheets with weakened basal texture by width-limited rolling, and the plasticity was significantly improved. Moreover, rare earth elements such as Y and Ce can cause double peak texture, which leads to a significant decrease in texture intensity [13, 14]. However, the complex process and expensive rare addition increase the consumption of test, which limit large-scale industrial usage of Mg alloys [4, 15]. Thereby, how to develop a new method with low cost is urgent. As is well known, with temperature increasing, the CRSS of non-basal slips is significantly reduced, which is beneficial to activate the non-basal slips [16]. The activation of non-basal slip further causes the grain orientation to be deflected, that is, the c-axis of the grain is no longer predominantly parallel to the normal direction [17]. Therefore, high temperature deformation can be used to weaken texture and reduce anisotropy without additional consumption. WANG et al [18] indicated that the in-plane anisotropy of ZK60 specimen was significantly weakened at 300 ℃ due to the activation of non-basal slip. WANG et al [19] concluded that at higher temperature, the increase of non-basal slips activity led to the decrease of normal direction anisotropy of AZ31 magnesium alloy. With the increase of temperature, the slips are activated due to the decrease of CRSS [20], and the difference of CRSS between basal slip and pyramidal slip decreased [13]. Meanwhile, there is no doubt that high-temperature rolling is also a potential effective process to obtain excellent magnesium alloy sheets with a weakened texture and great mechanical properties. However, rolling equipment at high temperature should be required and the energy cost as well as the loss of rolling mills is high. Therefore, to start the non-basal slips induced by thermal activation at higher temperatures with low cost is necessary and the process should be designed.

Thus, in this paper, a new two-step thermal pre-rolling technology is developed to achieve the non-basal slip on AZ31 Mg alloys sheet at higher temperatures. After that, normal rolling is followed to save costs. The microstructure evolution, texture behaviors and the mechanical response are investigated systematically. The detail mechanisms are discussed deeply.

2 Experimental procedure

Commercial AZ31 alloy sheets (Mg-3 wt.% Al-1 wt.% Zn) with a thickness of 4 mm were used as the investigated material in this work. The dimensions of the initial square samples are 80 mm×80 mm. Before rolling, all of the samples were annealed at 350 ℃ for 1 h, named as-received sample. The as-received samples were pre-rolled from 4 mm to 2.2 mm by 1 pass of 45% thickness reduction at different pre-rolling temperatures of 350 ℃ (350 ℃-PRI) and 450 ℃ (450 ℃-PRII). Then the pre-rolled samples were further rolled to 1 mm by 5 passes at 300 ℃, and the reduction of per pass was 15%. For comparison, other samples without pre-rolling were normal rolled by 5 passes at 300 ℃, and the reduction of per pass was 25%. During each pass, the samples were reheated for 5 min again to the rolling temperature. The samples after each pass are named the style of “temperature-pass”, such as 350 ℃-1P and 300 ℃-1P. Moreover, in order to obtain the performance of each pass sample, some samples were selected to anneal at 300 ℃ for 1 h in each pass, termed “temperature-passA” (e.g., 350 ℃-1PA, 300 ℃-5PA).

Dog-bone tensile sample with a gauge dimension of 12 mm×3 mm was machined from the samples annealed at 300 ℃ for 1 h. Then the uniaxial tensile tests were conducted at room temperature using a DNS200 electronic universal testing machine along rolling direction (RD), aligned 45°, and the transverse direction (TD). The strain rate was 1×10-3 s-1. The n-value was obtained from the uniform plastic deformation region of the true stress-strain curve. Selective tests were carried out at a permanent uniform strain without failure of 12% to measure the changes in sample dimensions (length and width) to determine the Lankford value (r-value).

The schematic diagram of rolling processes is given in Figure 1. After grinding, polishing and etching with a solution of picric acid, all samples annealed at 300 ℃ for 1 h were subjected to optical microscopy (OM; Leica DM2700 M). The microstructure evolutions were measured by OM and electron backscatter diffraction (EBSD) technologies. After grounding, polishing and electropolishing with commercial AC2 solution at a temperature of -20 ℃ using a voltage of 20 V for 90 s, EBSD analysis equipped with a scanning electron microscope (GEOLGSM-7800F FEG SEM) was used to identify the microstructure, and then an HKL channel 5 system was used to obtain the detailed analysis results.

Figure 1
Schematic diagram of rolling processes
pic

3 Results and discussion

3.1 Microstructure evolution before annealing

Figure 2 shows the OM and EBSD maps of the as-received sample. The microstructure is uniform and consists of equiaxed grains with an average size of 9.67 μm. No twins appear in the grain. As shown in Figure 2(b), most grains lay on basal plane, that is, the c-axis is basically parallel to normal direction (ND). Therefore, the initial AZ31 Mg alloy sheet presents a typical basal texture. Figure 2(c) gives the (0002) pole figure, which expresses a typical rolling basal texture with a relatively high texture intensity of 18.12.

Figure 2
Microstructure and EBSD maps of the as-received sample: (a) Optical micrograph; (b) Inverse pole figure (IPF); (c) (0002) pole figure; (d) Grain size map; (e) Misorientation angle map
pic

Optical micrographs after rolling are shown in Figure 3. It can be clearly seen from Figure 3(a) that a large number of twins appear in the grain of pre-rolling sample due to large reduction of 45%. In addition, recrystallized grains appear at twin and grain boundaries. However, the size of deformed grains decreases significantly, while it increases obviously on recrystallized grains with the increase of pre-rolling temperature. The volume fraction of dynamic recrystallized grains of the 450 ℃-PRII sample is larger than that of the 350 ℃-PRI sample, which is because the fine dynamic recrystallized grains begin to grow and consume the original grains. Compared with the 450 ℃-PRII sample, there are a large number of twins to coordinate plastic deformation in coarse grains of the 350 ℃-PRI sample, which indicates that the deformation mechanism is different at various pre-rolling temperatures. In the first pass rolling, the microstructure is composed of coarse grains with twins and recrystallized grains. However, the size of coarse grains decreases, and the grains become fine and equiaxial, especially in the 450 ℃-1P sample, as presented in Figures 3(d) and (e). After that, from the second pass to the fifth pass, the distribution of the microstructure becomes more uniform, and twins appear in almost every grain. For the comparison group, the microstructure evolution on directly rolling sample at 300 ℃ is different from that of pre-rolling samples, which is still not uniform and composed of fine recrystallized grains and coarsely deformed grains, as shown in Figures 3(c) and (f). To further investigate the deformation mechanism, EBSD is used for further analysis.

Figure 3
Optical micrographs of various pass thermal rolled AZ31 magnesium alloy sheets without annealing: (a) 350 ℃-PRI sample; (b) 450 ℃-PRII sample; (c) 300 ℃-1P sample; (d) 350 ℃-1P sample; (e) 450 ℃-1P sample; (f) 300 ℃-5P sample; (g) 350 ℃-5P sample; (h) 450 ℃-5P sample
pic

Figure 4 presents the EBSD maps of select samples without annealing. As shown in Figures 4(a) and (b), coarse grains are separated by twin boundaries, but the grain fragmentation is not complete, and some large grains still exist. In addition, the recrystallization phenomenon is obvious at the grain boundaries. As shown in Figures 4(d) and (e), the number of twins inside the coarse grains increases with the rolling process continues, dividing the large grains into small grains. It can be clearly seen in Figure 4(c) that many obvious recrystallized grains can be observed at the twin boundaries of the sample without pre-rolling. Moreover, the corresponding microstructure compositions of grains after dynamic recrystallization (DRXed grains), subgrains and deformed grains are also given in Figure 4. The size of the subgrains is larger than that of the DRXed grains and deformed grains. Dynamic recrystallization (DRX) leads to grain refinement as rolling progresses. As shown in Figures 4(d) and (e), the DRXed grains consist of a large number of fine DRXed grains, while the number is significantly reduced in Figures 4(a) and (b).

Figure 4
The EBSD maps of select samples without annealing: (a) 350 ℃-PRI sample; (b) 450 ℃-PRII sample; (c) 300 ℃-1P sample; (d) 350 ℃-1P sample; (e) 450 ℃-1P sample
pic
3.2 In-grain misorientation axes analysis

It is known to all that with the increase of temperature, the CRSS of non-basal slip systems reduces. Prismatic slip and pyramidal slip play an important role in the high temperature deformation process of magnesium alloy. Especially when the c-axis is parallel to the external force and in the hard orientation state, the pyramidal <a+c> slip can well coordinate the deformation of c-axis direction. In addition, previous studies have also proved that the activation of non-basal slip is closely related to dynamic recrystallization. HAN et al [21] pointed out that non-basal slips play an important role in the dynamic recrystallization. WEI et al [22] reported that the activation of slips is influenced by grain size. Therefore, the microstructure evolution may be related to the activation of non-basal slips. In order to find out the effect of non-basal slips on deformation, in-grain misorientation axes (IGMA) and texture component maps are used for further analysis.

By observing the intensity distribution of IGMA in a single grain, researchers can basically determine which slip system is activated. In theory, IGMA distribution can be divided into two types. When the IGMA intensity distribution is around <0001>, it is mainly caused by the activation of prismatic slip system [23]. However, when the intensity distribution is concentrated at <u v t 0>, which results from basal slip or pyramidal <a+c> slip [24]. Further, combined with the Schmid factor, the main activated slip system can be accurately determined. Figure 5 presents the IGMA distribution of numbered grains in Figure 4. As shown in Figure 5, several grains are selected from the denoised EBSD data to analyze their IGMA and the misorientation angle in this paper ranges from 2.5° to 5° [25, 26]. Although the IGMA intensity distribution of some grains spreads to (0001), the intensity distribution of selected grains is mainly concentrated around <u v t 0>. Moreover, the maximum intensity is greater than 2 and thus it is generally considered that there is a preferential IGMA. Therefore, it is speculated that basal slip and pyramidal <a+c> slip dominate the deformation.

Figure 5
The IGMA distribution of numbered grains in Figure 4
pic

In order to make a qualitative judgment on non-basal slip, Schmid factor is used for further analysis. The Schmid factor maps of as-received sample, 350 ℃-PRI sample and 450 ℃-PRII sample along RD and TD are shown is Figure 6. And the Schmid factor maps of 300 ℃-1P sample, 350 ℃-1P sample and 450 ℃-1P sample along RD and TD are shown in Figure 7. With the increase of rolling temperature, the difference of CRSS between basal slip and non-basal slip decreases. Therefore, the magnitude of Schmid factor is the main factor for the activation of slip systems. As shown in Figure 6, the Schmid factor values of basal slip of 350 ℃-PRI sample and 450 ℃-PRII sample along RD are 0.215 and 0.251, respectively. However, the values of pyramidal <a+c> slip of 350 ℃-PRⅠ sample and 450 ℃-PRII sample along RD are 0.395 and 0.375, respectively. It is clearly evident that the Schmid factor (SF) of pyramidal <a+c> slip is higher than that of basal slip. Moreover, the same trend is seen in TD. In the first pass, the same trend continues. In summary, the Schmid factor value of basal slip is lower than that of that of pyramidal <a+c> slip. Combined with the analysis of IGMA and Schmid factor, pyramidal <a+c> slip plays an important role in the deformation process.

Figure 6
The Schmid factor maps along RD (a-c) and TD (d-f): (a, d) As-received sample (AS); (b, e) 350 ℃-PRI sample; (c, f) 450 ℃-PRII sample
pic
Figure 7
The Schmid factor maps along RD and TD: (a, d) 300 ℃-1P sample; (b, e) 350 ℃-1P sample; (c, f) 450 ℃-1P sample
pic

Figure 8 shows the texture component maps of 350 ℃-1P sample. It is obvious that the basal texture and pyramidal texture dominate the overall texture component, while prismatic texture accounts for only a very small part, which indicates that pyramidal texture has a great influence on the deformation of pre-rolling sample. The activation of pyramidal <a+c> slip will weaken the texture and make the deformation easier. Therefore, the pre-rolling process is likely to promote the pyramidal slip and change the deformation mechanism.

Figure 8
The texture component maps of 350 ℃-1P sample
pic
3.3 Dynamic recrystallization mechanism
3.3.1 Discontinuous dynamic recrystallization

In the hot temperature rolling process, dynamic recrystallization behavior is complex, and different DRX mechanisms may play an irreplaceable role [27]. Continuous dynamic recrystallization (CDRX) and discontinuous dynamic recrystallization (DDRX) are two main recrystallization mechanisms, which may act independently or simultaneously [28, 29]. To clarify the recrystallization behavior during deformation, enlarged EBSD data were investigated.

Region II in Figure 4(c) is enlarged in Figure 9. In Figure 9(b), grains G1 and G12 are two parent grains of large size, and a chain of small recrystallized grains named G2-G11 can be clearly seen at the serrated grain boundaries. Figure 9(e) presents the misorientation profile along L2 marked in Figure 9(a), where the curve has two sharp peaks. Peak misorientation angles are 42.6° and 39.5°, respectively. Moreover, Figure 9(c) shows the (0002) pole figure of selected grains. From the pole figure, parent grains G1 and G10 are obviously deviated from the polar center about 38.5°. In general, serrated boundaries have high local stain gradients, which become a potential nucleation site for discontinuous dynamic recrystallization (DDRX) of grain boundary bulging [30, 31]. It can be clearly seen in Figure 9(b) that the parent grains show obvious serrated boundaries that generate DRXed grains, which is the feature of DDRX [31, 33]. Furthermore, DDRX helps to produce new grains with different orientation, which contributes to weak texture [32, 33]. Figure 9(d) presents the misorientation profile along L1. Neither point-to-origin misorientation nor point-to-point misorientation along L1 exceeds 10°, which indicates that the misorientation angle gradient is relatively low inside the parent grains.

Figure 9
Enlarged map of region II in Figure 4(c): (a) Enlarged IPF map; (b) Enlarged band contrast map; (c) Pole figure of grains marked in the IPF map; (d, e) Point-to-point and point-to-origin misorientation angle along L1 and L2
pic
3.3.2 Continuous dynamic recrystallization

The region III in Figure 4(c) is enlarged in Figure 10. As shown in Figure 10(b), the coarse grain is divided into several smaller grains by internally generated twins, which is marked with a green line, but possibly due to the large strain, recrystallized grains induced by twins are not fully detected in EBSD. However, observing the region corresponding to region III in the IPF map and DRXed grains map, the small recrystallized grains can be seen. Theoretically, the orientation angle of the pic double twin is 38.5° [34]. Figure 10(d) shows the misorientation angle along the L3 in Figure 10(a). The misorientation angle between G2 and G6 is 34.7°, indicating that grain G6 is formed by twin induction, although the misorientation changes slightly due to continuous deformation. Previous studies have also pointed out that some fine recrystallized grains can be seen within the origin grain due to subgrain rotation, shear bands and twins [28, 35]. In this way, grain orientation is changed and the recrystallized grains deviate from the surrounding parent grains as shown in Figure 10(c), which results in the texture weakening. PAN et al [24] and GUO et al [36] pointed out that subgrains were formed firstly in the twins, and then with low angle grain boundaries (LAGBs) growing to high angle grain boundaries (HAGBs), DRXed grains formed.

Figure 10
Enlarged map of region III in Figure 4(c); (a) Enlarged IPF map; (b) Enlarged band contrast map; (c) (0002) pole figure; (d) Point to point and point to origin misorientation angle along L3 in IPF
pic

Figure 11 shows the enlarged region I in Figure 4(a). As shown in Figure 11(a), a grain is divided into several subgrains by some LAGBs. However, it can be predicted that the misorientation angle between G2 and G4 will gradually increase and develop into recrystallized grains as the deformation continuous. Figures 11(d) and (e) present the misorientation angle map along L4 and L5. The misorientation angle of the point-to-point and point-to-origin curves does not exceed 15° either along grain interior L4 or across grain boundary L5. In addition, it is evident that the misorientation angle from point-to-origin is gradually increasing, which is characteristic of CDRX [32]. As can be seen from the pole figure (Figure 11(c)), there is no significant difference in the orientation of all grains.

Figure 11
Enlarged map of region I in Figure 4(a): (a) Enlarged IPF map; (b, c) Pole figure of grains marked in IPF map; (d, e) Point-to-point and point-to-origin misorientation angle along L4 and L5 in IPF
pic
3.4 Influence of double twin

It is well known that twinning can cause grain orientation changes so that the grains are in a favorable orientation to slip and further weaken the texture [37, 38]. Double twins are easily produced when the deformation temperature is higher than 300 ℃ [39, 40]. However, when the temperature is at 450 ℃, the effect of the double twin decreases. In the present work, due to the large strain of unannealed samples, it was difficult to distinguish double twins in the band contrast map. However, it can be clearly seen in the optical microscopy map of Figure 3(a) that the 350 ℃-PRI sample has a large number of double twins compared with the 450 ℃-PRII sample. Figure 12 shows the magnified view of region IV in Figure 4(d). Figure 12(d) shows the line profile for the microstructure angle along L6 in Figure 12(a). The red point-to-point line presents the misorientation angles across the double twin T, which are 39.6° and 36.5°, respectively. The existence of a pic double twin T colored green can be clearly seen in Figure 12(a). Moreover, the influence of the pic double twin on the texture distribution is also obvious. In Figure 12(c), the splitting of the texture along RD in the (0002) pole figure is caused by the double twin T. This is the most direct evidence that the texture split results from the appearance of the double twin.

Figure 12
Enlarged map of region IV in Figure 4(d): (a) Enlarged IPF map; (b) Enlarged band contrast map; (c) (0002) pole figure; (d) Point-to-point misorientation angle along L6 in IPF (pic double twin is colored in green)
pic

In summary, as shown in Figures 3(a), (b) and 11, CDRX mainly occurs in the coarse grains’ boundaries or twins generated within the coarse grains. However, due to the higher temperature of pre-rolling at 450 ℃, the fraction of twins is less than that of sample pre-rolled at 350 ℃. As the deformation continues, more twins are generated in the microstructure, especially in the 350 ℃ sample, thus the large grains are divided by twins. Therefore, the intersection of twins provides more nucleation site for DRX. As shown in Figures 3(d) and (e), DDRX induced at twin boundaries and CDRX generated inside twins are very common in subsequent normal rolling. As discussed in Section 3.2, the activation of non-basal slips has an important impact on the microstructure evolution. The activation of pyramidal <a+c> slip in the samples pre-rolled at 350 ℃ is better than that of the samples pre-rolled at 450 ℃, which leads to more obvious recrystallization in the samples pre-rolled at 350 ℃, while the samples pre-rolled at 450 ℃ may focus more on the growth of DRXed grains.

3.5 Texture evolution during rolling

Figure 13 shows the (0002) pole figures of the selected samples. It can be clearly seen in Figure 13(c) that the 300 ℃-1P sample exhibits a very high texture intensity of 20.64, which is higher than that of the as-received sample. Because the basal slip also plays a role in normal rolling. Therefore, it shows a typical basal texture, where the c-axis is parallel to ND, which results in poor plastic formability [37]. In contrast, from the pre-rolled pass to the first pass, the texture intensity of the 350 ℃-1P sample decreases significantly from 17.14 (Figure 13(a)) to 10.70 (Figure 13(d)) and that of the 450 ℃-1P sample decreases from 18.03 (Figure 13(b)) to 15.89 (Figure 13(e)). The former drops by as much as 37.6%, while the latter only shows 11.9%. After larger reductions of 45% and 15%, the texture intensities of the 350 ℃-1P and 450 ℃-1P samples are lower than that of the 300 ℃-1P sample, which is subjected to 25% reduction rolling. This suggests that pre-rolled process can dramatically weaken the basal texture, especially for the 350 ℃-1P sample, which is related to the activation of the pyramidal <a+c> slip determined by IGMA and Schmid factor. Texture weakening has been proven to be an effective method to improve the formability of magnesium alloys [38]. For this study, the texture distribution of pre-rolled samples is more dispersed than that of the normal rolled sample, i.e., the c-axis of grains inclined from ND to RD is more obvious. It is interesting that the 350 ℃-1P sample generated a double peak texture in RD, and the texture also shows an inclination from ND to TD, as shown in Figure 13(d). However, as the pre-rolling temperature increases to 450 ℃ in Figure 13(e), the 450 ℃-1P sample does not develop an obviously split texture and does not incline toward TD, which spreads to RD accompanied by a weak split. The texture evolution can be affected by rolling temperature, rolling reduction and initial texture [43, 44]. Both double and single peaks may occur in the rolled AZ31 sheet [44, 45]. HAN et al [46] also revealed that AZ31 sheet with equal channel angular rolling shows an unusual double peak texture in RD, and the formability is better than the conventional rolled sheet with strong basal texture. Previous studies have shown that the splitting of the basal poles in RD is associated with the double twin, while the inclined basal texture toward TD is related to the activity of prismatic slip [46-49]. In general, the strain in the width direction is provided by prismatic slip [24]. Therefore, it is speculated that the double peak texture is caused by double twins. However, the inclined texture toward the TD is associated with pyramidal slip, which is different from the previous studies. The intensity distribution of IGMA and texture component have shown that the inclined texture from ND to TD is mainly caused by pyramidal slips, especially pyramidal <a+c> slip rather than prismatic slip. Compared with 450 ℃-PRII sample, 350 ℃-PRI sample shows larger pyramidal <a+c> slip Schmid factor of 0.395, which indicates that 350 ℃-PRI sample is more likely to activate pyramidal <a+c> slip. Therefore, 350 ℃-1P sample shows significant inclined texture along TD. Additionally, it can be clearly seen in Figures 6(a) and (a) that basal slip in RD of 300 ℃-1P sample exhibits a smaller SF value of 0.179, while that of as-received sample is 0.203, which means that basal slip is still at work in the 300 ℃-1P sample, resulting in a higher texture intensity of 300 ℃-1P.

Figure 13
(0002) pole figures of select samples without annealing: (a) 350 ℃-PRI sample; (b) 450 ℃-PRII sample; (c) 300 ℃-1P sample; (d) 350 ℃-1P sample; (e) 450 ℃-1P sample
pic
3.6 Microstructure evolution after annealing

Figure 14 shows the IPF, misorientation angle maps and grain size distribution maps of the fifth pass samples annealed at 300 ℃ for 1 h. After annealing, the deformed grains and twins disappeared completely, and the microstructures present a typical basal texture and are composed of equiaxed recrystallized grains. The average grain sizes of the 300 ℃-5PA sample, 350 ℃-5PA sample and 450 ℃-5PA sample are 7.04, 4.31 and 5.70 μm, respectively. The 350 ℃-5PA sample shows the smallest average grain size of 4.31 μm. Moreover, although the average grain size of the 300 ℃-5PA sample is 7.04 μm, it can be seen clearly from the grain size distribution map that the microstructure of the 300 ℃-5PA sample shows a large volume of fine grains, resulting in a smaller average grain size in the final calculation. Different from the 300 ℃-5PA sample, the smaller average grain sizes of 350 ℃-5PA and 450 ℃-5PA are attributed to the fact that the grain sizes of those two samples are mostly distributed at approximately 5 μm. That is, the grain size distribution of the latter two is more uniform. The misorientation angles are characterized by recrystallized grain orientations of 35.75° and 35.00° for the 350 °C-5PA and 450 °C-5PA sample, respectively, while that of the 300 °C-5PA sample is 33.71°. Due to the activation of non-basal slip systems, the grain orientation is more dispersed, which makes the deformation easier and the accumulated energy increases. Furthermore, the grain orientation dispersed after recrystallization, and the 350 ℃-5PA sample is the most dispersed.

Figure 14
IPF maps, misorientation maps and grain size distribution maps of fifth pass samples annealed at 300 ℃ for 1 h: (a) 300 ℃-5PA sample; (b) 350 ℃-5PA sample; (c) 450 ℃-5PA sample
pic

Figure 15 shows the (0002), pic and pic pole figures of fifth pass samples annealed at 300 ℃ for 1 h. Although the non-basal slips contribute significantly to deformation, the texture is still basal texture. For the 300 ℃-5PA sample, the texture intensity is 12.80. The texture shows a normal rolling texture, i.e., the c-axis of grain inclines from ND to RD, as shown in Figure 15(a). Moreover, the double peak phenomenon is not detectable because the annealing process causes all twins to disappear. As shown in Figure 15(b), the texture distribution of 350 ℃-5PA changes, i.e., the c-axis of some grains slightly rotates away from RD to 45° due to annealing. Different from the 350 ℃-5PA sample mentioned above, the 450 ℃-5PA sample still shows a normal rolling texture with a c-axis inclined from ND to RD. After five passes, the thickness of all samples is approximately 1 mm. The texture intensities of the 350 ℃-5PA sample and 450 ℃-5PA are 10.97 and 10.61, respectively, which are less than that of the 300 ℃-5PA sample. This indicates that the grain orientation is more dispersed due to the influence of the annealing. Although there is a difference in texture intensity after annealing, the weakening of texture intensity by annealing is evident even after five rolling passes with a large reduction. Static recrystallization plays an important role in texture weaken.

Figure 15
(0002), pic, and pic pole figures of fifth pass samples annealed at 300 °C for 1 h: (a) 300 °C-5PA sample; (b) 350 °C-5PA sample; and (c) 450 °C-5PA sample
pic

4 Mechanical properties and anisotropy

Figure 16 illustrates the true stress-strain curve of sheets annealed at 300 °C for 1 h in the tensile directions of RD, 45° and TD. The curves of other pass samples are not shown due to similar trends. Regardless of the rolling temperature and reduction change, the flow stress increases with increasing strain until it reaches the maximum value and then decreases with increasing strain until fracture. Moreover, the ultimate tensile strength (UTS), yield stress (YS), r-value and n-value of the three fifth pass annealed samples are presented in Figure 17 for a better review. The YS in TD is the largest, but that in RD is the smallest, continuously decreasing from TD to RD. The regular changes in YS are reflected in the texture distribution. As shown in Figure 15, all the samples present an inclined texture from ND to RD. This indicates that the grain orientation is more dispersed in RD and the worst in TD.

Figure 16
The true stress-strain curve of selected sheets annealed at 300 ℃ for 1 h in the tensile direction of RD, 45° and TD: (a) 350 ℃-PRIA sample; (b) 450 ℃-PIIA sample; (c, f) 300 ℃-1PA sample and 300 ℃-5PA sample; (d, g) 350 ℃-1PA sample and 350 ℃-5PA sample; (e, h) 450 ℃-1PA sample and 450 ℃-5PA sample
pic
Figure 17
(a) YS and UTS, and (b) r-value and n-value of three fifth pass annealed samples in the direction of RD, 45° and TD
pic

Table 1 shows the average mechanical properties of three fifth pass annealed samples. For the three fifth pass annealed samples, the average fracture elongation is 17.31%, 18.83% and 16.81%, respectively, and the 350 ℃-5PA sample shows the largest average fracture elongation. The average YS and UTS of 350 ℃-5PA sample are higher than those of 300 ℃-5PA sample and 450 ℃-5PA sample. As shown in Figure 14, the smallest grain size is obtained in the 350 ℃-5PA sample. It is generally thought that the plasticity and strength of metal materials cannot be improved at the same time, but fine grain strengthening is one of the few ways to increase both strength and plasticity [50-52]. According to the Hall-Petch formula, the smaller the grain size is, the greater the yield strength of the material is. In addition, when the material with fine grains is affected by an external force, the plastic deformation can be dispersed in more grains and is uniform, which results in the improvement of plasticity. These results show that the mechanical properties of the 350 ℃-5PA is the best. The sample subjected to 350 ℃ pre-rolling shows a smaller pic and a larger pic. A smaller r-value indicates that a reduction in thickness is easier during plastic deformation, resulting in less normal anisotropy in the sheet. A larger n-value means a lower strain concentration sensitivity during necking, which is beneficial to increase uniform elongation. Therefore, the fracture elongation of the 350 ℃-5PA sample is the largest. For the 350 ℃-5PA sample, except for RD, where the r-value and n-value fluctuated slightly, the r-value is the minimum and the n-value is the maximum in other directions. As shown in Figure 17(b), the largest differences of r-value and n-value are found in the 45°, which decreases from 5.52 to 2.79 and increases from 0.25 to 0.37, respectively. This is related to the evolution of texture, as shown in Figure 15. As shown in Figure 15(b), the 350 ℃-5PA sample shows a slightly rotating texture from RD to 45°, so that the grain orientation near 45° is more scattered, which leads to the weakness of anisotropy.

Table 1
Average mechanical properties of the fifth-pass samples
SampleYS/MPaUTS/MPapic/%picpic
300 ℃-5P186.90307.1617.314.380.26
350 ℃-5P194.12312.8718.832.470.32
450 ℃-5P175.70291.9416.812.700.27
展开更多

Figure 18 shows the ΔYS and Δr values of three fifth pass samples. As shown in Figure 18(a), the ΔYS of 300 ℃-5PA sample, 350 ℃-5PA sample and 450 ℃-5PA sample are 5.77, 3.75 and 6.20, respectively. After five passes, the ΔYS of the 350 ℃-5PA sample is the smallest, indicating that the plane anisotropy reaches the minimum compared with the 300 ℃-5PA sample and 450 ℃-5PA sample. A similar phenomenon is also reflected in Δr. As shown in Figure 18(b), the Δr of three fifth pass annealed samples are 2.28, 0.65 and 1.14, respectively. 300 ℃-5PA sample presents a larger Δr, while after the pro-rolling process, the Δr of 350 ℃-5PA sample and 450 ℃-5PA sample decreases significantly, especially for the 350 ℃-5PA sample, which indicates that pre-rolling process is an effective way to reduce the anisotropy of the AZ31 magnesium sheet. SHI et al [53] pointed out that prismatic slip is the main reason for the high anisotropy of magnesium alloy sheets, while the activity of pyramidal <a+c> slip. The r-value of sheet decreases significantly, that is, the anisotropy weakens significantly [54]. In this study, the activity of pyramidal <a+c> slip promotes the occurrence of DDRX, dispersing the grain orientation, weakening the basal texture and further weakening the anisotropy. Under the influence of the deformation condition, the activity of the pyramidal <a+c> slip of 350 ℃ pre-rolling sample is the most obvious, which results in weakened anisotropy.

Figure 18
(a) ΔYS and (b) Δr value of the fifth-pass samples
pic

5 Conclusions

Pre-rolling is performed at 350 and 450 ℃ to investigate the anisotropy weakening of AZ31 Mg alloy sheet. After the rolling process, the anisotropy of pre-rolling samples is compared with the sample without pre-rolling. The main conclusions are summarized below.

The high temperature pre-rolling process activates non-basal slip and disperses grain orientation, which results in dynamic recrystallization more active. In addition, due to the influence of twins, the texture intensity decreases significantly and the texture distribution disperses obviously.

The texture intensities of the 350 ℃-5PA and 450 ℃-5PA samples are 10.97 and 10.61, respectively, which are lower than that of the 300 ℃-5PA sample (12.80). The 350 ℃-5PA sample shows a unique texture, where the c-axis is slightly rotated from RD to 45°, which is beneficial to weaken anisotropy.

Compared with the sample without pre-rolling and the 450 ℃-5PA sample, the 350 ℃-5PA sample shows the smallest Δr of 0.65, which indicates that the 350 ℃-5PA sample exhibits less anisotropy in the sheet. In summary, the 350 ℃ pre-rolling process is an effective method to obtain magnesium alloys with a weakened anisotropy.

References
1GAO Ju-ming, WANG Dan, ZHUANG Dong-dong, et al.

Effect of laser power on microstructure and mechanical properties of laser heat conduction lap welded joint between AZ31B magnesium alloy and DP780 galvanized steel

[J]. Journal of Central South University, 2022, 29(10): 3463-3475. DOI: 10.1007/s11771-022-5164-9.
百度学术谷歌学术
2YIN Y L, YU H L, WANG H M, et al.

Tribological behaviors of AZ91D magnesium alloy under the lubrication of oil suspended synthetic magnesium silicate hydroxide nanotubes

[J]. Journal of Magnesium and Alloys, 2024. DOI: 10.1016/j.jma.2024.01.007.
百度学术谷歌学术
3CHU Chen-liang, WU Xiao-quan, QIU Shui-cai, et al.

Microstructure and Gd-rich phase evolution of as-cast AZ31-xGd magnesium alloys during semi-solid isothermal heat treatment

[J]. Journal of Central South University, 2021, 28(1): 1-15. DOI: 10.1007/s11771-020-4504-x.
百度学术谷歌学术
4JI Ya-feng, DUAN Jin-rui, YUAN Hao, et al.

Effect of variable thickness cross rolling on edge crack and microstructure gradient of AZ31 magnesium alloy

[J]. Journal of Central South University, 2022, 29(4): 1124-1132. DOI: 10.1007/s11771-022-4973-1.
百度学术谷歌学术
5ZHANG Yun, JIANG Hai-tao, KANG Qiang, et al.

Microstructure evolution and mechanical property of Mg-3Al alloys with addition of Ca and Gd during rolling and annealing process

[J]. Journal of Magnesium and Alloys, 2020, 8(3): 769-779. DOI: 10.1016/j.jma.2019.11.015.
百度学术谷歌学术
6ZHANG Wen-cong, LIU Xin-tong, MA Jun-fei, et al.

Evolution of microstructure and mechanical properties of ZK60 magnesium alloy processed by asymmetric lowered-temperature rolling

[J]. Transactions of Nonferrous Metals Society of China, 2022, 32(9): 2877-2888. DOI: 10.1016/S1003-6326(22)65990-9.
百度学术谷歌学术
7OYINBO S T, SINGHANEKA S, MATSUMOTO R.

Exploring the basal/prismatic slip transfer at grain boundaries in magnesium: A molecular dynamic simulation

[J]. Vacuum, 2023, 212: 111995. DOI: 10.1016/j.vacuum. 2023.111995.
百度学术谷歌学术
8ZHANG Kai, SHAO Zhu-tao, ROBSON J, et al.

Achieving high ductility and strength in magnesium alloy through cryogenic-hot forming

[J]. Journal of Magnesium and Alloys, 2023, 11(9): 3130-3140. DOI: 10.1016/j.jma.2023.09.002.
百度学术谷歌学术
9XIE Zhen-dong, GUAN Yan-jin, YU Xiao-hui, et al.

Effects of ultrasonic vibration on performance and microstructure of AZ31 magnesium alloy under tensile deformation

[J]. Journal of Central South University, 2018, 25(7): 1545-1559. DOI: 10.1007/s11771-018-3847-z.
百度学术谷歌学术
10ZHANG Wen-cong, LIU Yu-xuan, MA Jun-fei, et al.

Effects of bimodal basal texture on bending stress and microstructure evolution of Mg-2.6Nd-0.55Zn-0.5Zr alloys

[J]. Transactions of Nonferrous Metals Society of China, 2023, 33(3): 714-727. DOI: 10.1016/S1003-6326(22)66140-5.
百度学术谷歌学术
11ZHANG Dong-dong, LIU Chu-ming, WAN Ying-chun, et al.

Microstructure and anisotropy of mechanical properties in ring rolled AZ80-Ag alloy

[J]. Journal of Central South University, 2021, 28(5): 1316-1323. DOI: 10.1007/s11771-021-4699-5.
百度学术谷歌学术
12TIAN Jing, LU Hui-hu, ZHANG Wang-gang, et al.

An effective rolling process of magnesium alloys for suppressing edge cracks: Width-limited rolling

[J]. Journal of Magnesium and Alloys, 2022, 10(8): 2193-2207. DOI: 10.1016/j.jma. 2021.01.007.
百度学术谷歌学术
13WANG Li-fei, LI Yong-qiao, ZHANG Hua, et al.

Review: Achieving enhanced plasticity of magnesium alloys below recrystallization temperature through various texture control methods

[J]. Journal of Materials Research and Technology, 2020, 9(6): 12604-12625. DOI: 10.1016/j.jmrt.2020.09.002.
百度学术谷歌学术
14ROKHLIN L L.

The regularities in constitution of the ternary phase diagrams of magnesium alloys with two rare-earth metals belonging to different groups (yttrium group or cerium group)

[J]. Journal of Phase Equilibria and Diffusion, 2020, 41(4): 532-537. DOI: 10.1007/s11669-020-00817-1.
百度学术谷歌学术
15LI Z T, QIAO X G, XU C, et al.

Enhanced strength by precipitate modification in wrought Mg-Al-Ca alloy with trace Mn addition

[J]. Journal of Alloys and Compounds, 2020, 836: 154689. DOI: 10.1016/j.jallcom.2020.154689.
百度学术谷歌学术
16NING Fang-kun, LE Qi-chi, YAN Jia-shi, et al.

Crimp feasibility of AZ31 magnesium alloy wide plate at warm temperatures responding to asymmetry

[J]. Journal of Magnesium and Alloys, 2022, 10(3): 797-810. DOI: 10.1016/j.jma.2020.08.020.
百度学术谷歌学术
17SHI D F, CEPEDA-JIMÉNEZ C M, PÉREZ-PRADO M T.

The relation between ductility at high temperature and solid solution in Mg alloys

[J]. Journal of Magnesium and Alloys, 2022, 10(1): 224-238. DOI: 10.1016/j.jma.2021.09.024.
百度学术谷歌学术
18WANG Wen-ke, CHEN Wen-zhen, ZHANG Wen-cong, et al.

Weakened anisotropy of mechanical properties in rolled ZK60 magnesium alloy sheets with elevated deformation temperature

[J]. Journal of Materials Science & Technology, 2018, 34(11): 2042-2050. DOI: 10.1016/j.jmst.2018.02.019.
百度学术谷歌学术
19WANG Li-fei, PAN Xiao-huan, ZHU Xing-xiao, et al.

Effect of two-step increased temperature thermal rolling on anisotropy and stretch formability of AZ31 magnesium alloy sheets

[J]. Transactions of Nonferrous Metals Society of China, 2023, 33(4): 1066-1085. DOI: 10.1016/S1003-6326(23)66166-7.
百度学术谷歌学术
20GONG Hai, HE Yong-biao, ZHANG Tao, et al.

Modified constitutive behavior model of Mg-10Gd-3Y-0.4Zr alloy during high-temperature deformation process

[J]. Journal of Central South University, 2023, 30(8): 2458-2471. DOI: 10.1007/s11771-023-5409-2.
百度学术谷歌学术
21HAN Lu-yi, YU Ya-nan, WEI De-jin, et al.

The synergistic and interactive effects of slip systems and dynamic recrystallization on the weakening basal texture of Mg-Y-Nd-Zr-Gd magnesium alloy

[J]. Materials & Design, 2024, 237: 112583. DOI: 10.1016/j.matdes.2023.112583.
百度学术谷歌学术
22WEI Kang, HU Rong, YIN Dong-di, et al.

Grain size effect on tensile properties and slip systems of pure magnesium

[J]. Acta Materialia, 2021, 206: 116604. DOI: 10.1016/j.actamat. 2020.116604.
百度学术谷歌学术
23ZHU Yu-zhi, HOU De-wen, LI Qi-zhen.

Quasi in situ EBSD analysis of twinning-detwinning and slip behaviors in textured AZ31 magnesium alloy subjected to compressive-tensile loading

[J]. Journal of Magnesium and Alloys, 2022, 10(4): 956-964. DOI: 10.1016/j.jma.2021.08.025.
百度学术谷歌学术
24PAN Xiao-huan, WANG Li-fei, LU Peng-bin, et al.

Unveiling the planar deformation mechanisms for improved formability in pre-twinned AZ31 Mg alloy sheet at warm temperature

[J]. Journal of Magnesium and Alloys, 2023, 11(12): 4659-4678. DOI: 10.1016/j.jma.2022.11.010.
百度学术谷歌学术
25SUH B C, KIM J H, BAE J H, et al.

Effect of Sn addition on the microstructure and deformation behavior of Mg-3Al alloy

[J]. Acta Materialia, 2017, 124: 268-279. DOI: 10. 1016/j.actamat.2016.11.020.
百度学术谷歌学术
26WU Wei, WANG Li-fei, HUANG Guang-sheng, et al.

Effect of multi-pass continuous screw twist extrusion process on microstructure evolution, texture, and mechanical properties of AZ31 magnesium alloy

[J]. Materials Today Communications, 2023, 34: 105508. DOI: 10.1016/j.mtcomm.2023.105508.
百度学术谷歌学术
27MIRZADEH H.

Grain refinement of magnesium alloys by dynamic recrystallization (DRX): A review

[J]. Journal of Materials Research and Technology, 2023, 25: 7050-7077. DOI: 10.1016/j.jmrt.2023.07.150.
百度学术谷歌学术
28LI Zhao-can, WU Guo-qin, YU Jian-min, et al.

Influence of strain rate on grain refinement and texture evolution under complex shear stress conditions

[J]. Journal of Magnesium and Alloys, 2023, 11(7): 2558-2584. DOI: 10.1016/j.jma. 2022.11.020.
百度学术谷歌学术
29HU Jian-liang, WU Xiu-jiang, BO Hong, et al.

Dislocation density model and microstructure of 7A85 aluminum alloy during thermal deformation

[J]. Journal of Central South University, 2021, 28(10): 2999-3007. DOI: 10.1007/s11771-021-4832-5.
百度学术谷歌学术
30HAO Mei-juan, CHENG Wei-li, WANG Li-fei, et al.

Texture evolution induced by twinning and dynamic recrystallization in dilute Mg-1Sn-1Zn-1Al alloy during hot compression

[J]. Journal of Magnesium and Alloys, 2020, 8(3): 899-909. DOI: 10.1016/j.jma.2019.10.002.
百度学术谷歌学术
31SUN Dong, JIANG Shu-yong, ZHANG Yan-qiu, et al.

Discontinuous dynamic recrystallization of TiNb alloys: Experiment and cellular automaton simulation

[J]. Journal of Central South University, 2023, 30(9): 2890-2905. DOI: 10.1007/s11771-023-5430-5.
百度学术谷歌学术
32DU Hua-qiu, LI Feng, HUO Peng-da, et al.

Microstructure evolution and ductility improvement mechanisms of magnesium alloy in interactive alternating forward extrusion

[J]. Transactions of Nonferrous Metals Society of China, 2022, 32(8): 2557-2568. DOI: 10.1016/S1003-6326(22)659 65-X.
百度学术谷歌学术
33LIANG Min-jie, ZHENG Jie, LIU Huan, et al.

Microstructure and mechanical properties of AZ31 alloy prepared by cyclic expansion extrusion with asymmetrical extrusion cavity

[J]. Transactions of Nonferrous Metals Society of China, 2022, 32(1): 122-133. DOI: 10.1016/S10 03-6326(21)65782-5.
百度学术谷歌学术
34LI Hao, ZENG Qing-hui, YANG Peng-fei, et al.

Towards understanding twinning behavior near fracture surface in magnesium

[J]. Journal of Materials Science & Technology, 2020, 43: 230-237. DOI: 10.1016/j.jmst.2020.01.007.
百度学术谷歌学术
35XIE Bing-chao, ZHANG Bao-yun, NING Yong-quan, et al.

Mechanisms of DRX nucleation with grain boundary bulging and subgrain rotation during the hot working of nickel-based superalloys with columnar grains

[J]. Journal of Alloys and Compounds, 2019, 786: 636-647. DOI: 10.1016/j.jallcom. 2019.01.334.
百度学术谷歌学术
36GUO Peng-cheng, LIU Xiao, ZHU Bi-wu, et al.

The microstructure evolution and deformation mechanism in a casting AM80 magnesium alloy under ultra-high strain rate loading

[J]. Journal of Magnesium and Alloys, 2022, 10(11): 3205-3216. DOI: 10.1016/j.jma.2021.07.032.
百度学术谷歌学术
37SONG Bo, WANG Meng, SHI Ruo-lan, et al.

Promoting hybrid twins structure to reduce yield asymmetry of rolled AZ31 plates by combining side-rolling and torsion

[J]. Journal of Magnesium and Alloys, 2023, 11(6): 2096-2105. DOI: 10.1016/j.jma.2021.08.012.
百度学术谷歌学术
38HU Zhi-qing, GUO Chao-fan, LI Hong-mei.

Influence of AZ31 sheet treated by cryogenic on punch shearing

[J]. Journal of Central South University, 2019, 26(6): 1582-1591. DOI: 10.1007/s11771-019-4114-7.
百度学术谷歌学术
39CHAPUIS A, DRIVER J H.

Temperature dependency of slip and twinning in plane strain compressed magnesium single crystals

[J]. Acta Materialia, 2011, 59(5): 1986-1994. DOI: 10.1016/j.actamat.2010.11.064.
百度学术谷歌学术
40LIU Yu-xuan, LI Yang-xin, ZHU Qing-chun, et al.

Twin recrystallization mechanisms in a high strain rate compressed Mg-Zn alloy

[J]. Journal of Magnesium and Alloys, 2021, 9(2): 499-504. DOI: 10.1016/j.jma.2019.12.011.
百度学术谷歌学术
41YIN Gui-mei, FENG Zhi-yuan, ZHAO Zi-long, et al.

Microstructure and creep properties of Al/Mg/Al composite clad plates

[J]. Journal of Central South University, 2023, 30(9): 2867-2877. DOI: 10.1007/s11771-023-5437-y.
百度学术谷歌学术
42LIU Di, LIU Zu-yan, WANG Er-de.

Improving single pass reduction during cold rolling by controlling initial texture of AZ31 magnesium alloy sheet

[J]. Transactions of Nonferrous Metals Society of China, 2018, 28(2): 244-250. DOI: 10.1016/S1003-6326(18)64657-6.
百度学术谷歌学术
43MALIK A, WANG Yang-wei, NAZEER F, et al.

Effect of pre-straining on twinning, texture and mechanical behavior of magnesium alloys—A review

[J]. Journal of Materials Research and Technology, 2020, 9(6): 14478-14499. DOI: 10.1016/j.jmrt.2020.10.023.
百度学术谷歌学术
44SONG Deng-hui, ZHOU Tao, TU Jian, et al.

Improved stretch formability of AZ31 sheet via texture control by introducing a continuous bending channel into equal channel angular rolling

[J]. Journal of Materials Processing Technology, 2018, 259: 380-386. DOI: 10.1016/j.jmatprotec. 2018.05.012.
百度学术谷歌学术
45GUO Fei, PEI Ri-sheng, JIANG Lu-yao, et al.

The role of recrystallization and grain growth in optimizing the sheet texture of magnesium alloys with calcium addition during annealing

[J]. Journal of Magnesium and Alloys, 2020, 8(1): 252-268. DOI: 10.1016/j.jma.2019.07.010.
百度学术谷歌学术
46HAN Xiu-zhu, HU Li, JIA Dong-yong, et al.

Role of unusual double-peak texture in significantly enhancing cold rolling formability of AZ31 magnesium alloy sheet

[J]. Transactions of Nonferrous Metals Society of China, 2023, 33(8): 2351-2364. DOI: 10.1016/S1003-6326(23)66264-8.
百度学术谷歌学术
47LIU Bo-shu, XUE Li-kai, LI Rong-guang, et al.

Plasticity damage behavior caused by compression twins and double twins in rolled WE43 magnesium alloys

[J]. Materials Letters, 2023, 350: 134877. DOI: 10.1016/j.matlet.2023. 134877.
百度学术谷歌学术
48ELSAYED F R, SASAKI T T, MENDIS C L, et al.

Compositional optimization of Mg-Sn-Al alloys for higher age hardening response

[J]. Materials Science and Engineering A, 2013, 566: 22-29. DOI: 10.1016/j.msea. 2012.12.041.
百度学术谷歌学术
49NARAYAN J.

Recent progress in thin film epitaxy across the misfit scale (2011 Acta Gold Medal Paper)

[J]. Acta Materialia, 2013, 61(8): 2703-2724. DOI: 10.1016/j.actamat. 2012.09.070.
百度学术谷歌学术
50WEN Kai, LI Xi-wu, XIONG Bai-qing, et al.

Influence of minor Sc additions on grain refinement and microstructure characteristics of a high Zn-containing Al-Zn-Mg-Cu-Zr alloy

[J]. Journal of Central South University, 2022, 29(3): 780-794. DOI: 10.1007/s11771-022-4979-8.
百度学术谷歌学术
51GUI Yun-wei, OUYANG Ling-xiao, CUI Yu-jie, et al.

Grain refinement and weak-textured structures based on the dynamic recrystallization of Mg-9.80Gd-3.78Y-1.12Sm-0.48Zr alloy

[J]. Journal of Magnesium and Alloys, 2021, 9(2): 456-466. DOI: 10.1016/j.jma.2020.06.001.
百度学术谷歌学术
52FAN Z, GAO F, WANG Y, et al.

Grain refinement of Mg-alloys by native MgO particles: An overview

[J]. Journal of Magnesium and Alloys, 2022, 10(11): 2919-2945. DOI: 10.1016/j.jma.2022.10.006.
百度学术谷歌学术
53SHI Bao-dong, YANG Chong, PENG Yan, et al.

Anisotropy of wrought magnesium alloys: A focused overview

[J]. Journal of Magnesium and Alloys, 2022, 10(6): 1476-1510. DOI: 10.1016/j.jma.2022.03.006.
百度学术谷歌学术
54PAN Xiao-huan, WANG Li-fei, XUE Liang-liang, et al.

Dynamic recrystallization, twinning behaviors and mechanical response of pre-twinned AZ31 Mg alloy sheet along various strain paths at warm temperature

[J]. Journal of Materials Research and Technology, 2022, 19: 1627-1649. DOI: 10.1016/j.jmrt.2022.05.164.
百度学术谷歌学术
注释

YANG Chao-yang, WANG Li-fei, XUE Liang-liang, HUANG Qiu-yan Huan, XIA Da-biao, FU Xin-wei, SONG Bo, ZHENG Liu-wei, WANG Hong-xia and KWANG Seon-Shin declare that they have no conflict of interest.

YANG Chao-yang, WANG Li-fei, XUE Liang-liang, HUANG Qiu-yan, XIA Da-biao, FU Xin-wei, SONG Bo, ZHENG Liu-wei, WANG Hong-xia, KWANG Seon-Shin. Weakened in-plane anisotropy of AZ31 magnesium alloy sheet induced by pre-enhanced non-basal slips during hot rolling [J]. Journal of Central South University, 2025, 32(3): 706-726. DOI: https://doi.org/10.1007/s11771-025-5856-z.

杨朝阳,王利飞,薛亮亮等.热轧过程促进非基面滑移开动弱化AZ31镁合金板材平面各向异性研究[J].中南大学学报(英文版),2025,32(3):706-726.