Next Article in Journal
The Corrosion Behavior of X100 Pipeline Steel in a Sodium Chloride Solution Containing Magnesium and Calcium
Next Article in Special Issue
Effect of Deformation Conditions on Strain-Induced Precipitation of 7Mo Super-Austenitic Stainless Steel
Previous Article in Journal
The Removal of Inclusions with Different Diameters in Tundish by Channel Induction Heating: A Numerical Simulation Study
Previous Article in Special Issue
Effects of V Addition on the Deformation Mechanism and Mechanical Properties of Non-Equiatomic CoCrNi Medium-Entropy Alloys
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Asymmetric Extrusion Technology of Mg Alloy: A Review

1
School of Metallurgy and Material Engineering, Chongqing University of Science and Technology, Chongqing 401331, China
2
National Engineering Research Center for Magnesium Alloys, Chongqing University, Chongqing 400044, China
3
Department of Materials Science and Engineering, The Ohio State University, Columbus, OH 43210, USA
*
Authors to whom correspondence should be addressed.
Materials 2023, 16(15), 5255; https://doi.org/10.3390/ma16155255
Submission received: 19 June 2023 / Revised: 19 July 2023 / Accepted: 24 July 2023 / Published: 26 July 2023

Abstract

:
Magnesium (Mg) alloy is a widely used lightweight metal structural material due to its high specific strength and stiffness, excellent damping performance, and recyclability. Wrought Mg alloys are particularly favored in fields such as aerospace, transportation, and biomedical stents. However, most wrought Mg alloys with a hexagonal close-packed (HCP) crystal structure lack sufficient independent slip systems to meet the von Mises criterion for uniform plastic deformation at room temperature. This can result in the formation of a strong basal texture during plastic deformation and poor room temperature plastic formability. Enhancing the room temperature forming performance is therefore a crucial challenge that needs to be addressed in order to expand the application of Mg alloy sheets. Our research group has comprehensively summarized significant work and the latest research progress in improving the room temperature forming of Mg alloy sheets via extrusion technology in recent years. Specifically, we have developed a new type of asymmetric extrusion technology that combines material structure evolution, mechanical properties, and forming behavior analysis. We have elucidated the extrusion process characteristics, texture control mechanism, and forming properties of Mg alloy sheets through plastic deformation mechanisms, mold design, and finite element numerical simulation. The findings of our study present an innovative extrusion technology for the fabrication of highly formable Mg alloy sheets, which can be utilized in various applications.

1. Introduction

With the rapid advancement of technology, there is an increasing demand for lightweight and high-strength structural materials in key sectors such as aviation, aerospace, transportation, and high-end equipment manufacturing [1,2,3,4]. Magnesium (Mg) alloy has emerged as one of the most competitive lightweight metal structural materials [5,6,7,8]. In particular, the demand for high-performance Mg alloys in major lightweight projects becomes even more critical, as it holds strategic significance for achieving structural lightweight, energy savings, emission reduction, and safe service [9,10,11]. However, the large critical shear stress difference required for the activation of basal and non-basal slip systems in Mg alloys results in the main slip system being basal slip during plastic deformation [12,13,14]. Conventionally deformed Mg alloys by plastic processing have strong basal texture and anisotropy [15,16,17,18]. The poor room temperature formability and difficult processing and forming of conventionally processed deformed Mg alloys significantly limit their large-scale application and development [19,20,21,22,23]. Therefore, improving the room temperature forming performance of Mg alloy sheets remains one of the crucial problems that urgently needs to be addressed.
In recent years, extensive research has been conducted on the poor room temperature formability of Mg alloys, with texture control emerging as a current research hotspot [22,23,24,25,26,27]. Currently, texture control technologies primarily focus on two aspects: trace alloying element addition and plastic deformation processing. The addition of rare earth elements such as Nd, Gd, and Y has been shown to weaken basal texture [28,29,30,31]. However, precise control of the rare earth element content is necessary, as an excessive addition may lead to the formation of second-phase particles that are not conducive to subsequent plastic forming processes. Moreover, rare earth elements are expensive [32,33,34]. Traditional plastic processing methods such as hot extrusion, warm rolling, and cold rolling are commonly used in the processing and preparation of Mg alloy sheet [35,36,37]. These methods result in most crystallites of magnesium alloy sheets being almost parallel to the normal direction of the sheet, exhibiting lower ductility and formability [38,39,40,41].
Extrusion is a commonly used processing method for preparing Mg alloy sheets [42,43,44]. During plastic deformation, Mg alloys are strongly influenced by external stresses, leading to directional flow and coordinated rotation of the grains relative to the axis of the external force, resulting in the formation of a deformation texture. Changes in external stresses can cause shifts in crystal rotation trends, leading to corresponding changes in the deformation texture. By utilizing specialized extrusion processes, it is possible to control the temperature and stress states during deformation. This can eliminate strong basal plane textures that form due to compressive deformations in the thickness direction of the sheet and thereby adjust and control the texture of the Mg alloy sheet [45,46,47,48]. This approach has become an important means of preparing high-performance Mg alloy sheets and improving their subsequent forming abilities.
During the extrusion process, a heated alloy ingot is loaded into the extrusion cylinder of the machine and subjected to strong triaxial compression stress. The resulting Mg alloy sheet has a certain width and thickness after being passed through a specific rectangular die. Conventional extrusion (CE) processes involve symmetrical extrusion forces, leading to a strong basal texture and isotropy of the Mg alloy sheet. However, a new type of asymmetric extrusion for Mg alloy sheets involves constructing asymmetrical internal geometries within the extrusion die to maintain asymmetrical stress and strain during extrusion [49,50,51]. This can increase additional shear strain, refine grain size, overcome the dead zone phenomenon, and improve the smoothness of metal flow and flow properties of the metal extrusion process. The flow rate gradient and strain gradient formed during this process cause the c-axis orientation of the sheet grains to tilt along the extrusion direction, weakening the basal texture of the sheet and improving overall mechanical properties [52,53,54].
The author’s team has developed various new types of asymmetric geometric extrusion dies. This was mainly achieved by introducing different gradient strains from the thickness direction (normal direction) and the transverse direction of the extruded sheets, as well as changing the flow rate, strain, and other parameters of the extruded Mg alloy sheet. Ultimately, this resulted in the regulation of crystal orientation. Different extrusion processes were employed, including asymmetric extrusion (ASE) [55,56,57], differential speed extrusion (DSE) [58], normal gradient extrusion (NGE) [59], transverse gradient extrusion (TGE) [60], asymmetric porthole die extrusion (APE) [61], asymmetric material composition extrusion [62], and asymmetric curve extrusion (ACE) [63]. Currently, AZ31 alloy in the Mg-Al series is characterized by its extensive usage and cost-effectiveness. A large amount of research has been focused on the properties and microstructural regulation of AZ31. Table 1 summarizes the mechanical properties of Mg alloy sheets processed by different extrusion technologies. FE, r, and n represent fracture elongation, the Lankford value, and the strain hardening exponent value, respectively. It is evident that asymmetric extrusion technology can reduce anisotropy, introduce shear deformation to facilitate grain deviation, promote the activity of basal <a> slip initiation [64,65,66], and improve the plastic deformation ability of magnesium alloy while enhancing processing efficiency for Mg alloy sheet preparation.

2. Processing Extrusion Technologies of Mg Alloy

2.1. Normal-Direction Asymmetric Extrusion Technology of Mg Alloy

Normal-direction asymmetric extrusion is a promising technology used for processing Mg alloy sheets. This technique involves the use of a heat-treated Mg alloy ingot processed in an extruder equipped with an asymmetric extrusion platform and die. During the extrusion process, the sheet undergoes strong triaxial compressive stress and is then extruded through the die to obtain a Mg alloy sheet with a specific width and thickness. Compared with traditional symmetrical extrusion, thick-directional asymmetric extrusion introduces additional shear strain, resulting in a finer grain size and improved smoothness of metal flow. Consequently, it enhances the flow ability of the metal extrusion process. By adjusting the contact distance and angle between the working belt of the die and the upper and lower surfaces of the sheet, a non-symmetric shear strain gradient along the thickness direction of the sheet (parallel to the thickness plane of the sheet) can be formed. Shear deformation along the direction parallel to the sheet’s thickness will induce grain orientation with localized strain, which can significantly improve the basal texture of Mg alloy thin sheets [49,68]. Thus, the c-axis orientation of the sheet grains tilts along the direction of extrusion, weakening the basal texture of the sheet and thus improving the overall mechanical performance of the extruded sheet.

2.1.1. Asymmetric Extrusion (ASE)

Figure 1 shows the (0002) pole figures and corresponding EBSD analysis of CE and ASE extruded sheets. The CE sheet exhibits a strong (0002) basal plane texture with relatively uniform organization, while the ASE sheet displays a weakened basal plane texture but uneven organization in the thickness direction. As presented in Figure 2b, numerous small dynamic recrystallization (DRX) grains can be observed around the elongated grains, and the ED deviation of the basal plane texture axis is approximately 12°. Coarse grains are elongated and then deviate from the basal orientation.
The CE sheet undergoes a strain that compresses its thickness while elongating it along the extrusion direction. In contrast, the Mg alloy sheet during ASE processing also experiences shear stress due to the different velocities of the upper and lower parts, resulting in shearing deformation in the sheet thickness direction, with the slow side moving rearward and the fast side moving forward. As a result of the significant extrusion deformation and a high extrusion ratio of approximately 100:1, the Mg alloy sheet has a high stored energy, and the driving force for recrystallization is strong, leading to a high nucleation and growth rate. However, when the deformation amount is significant, the nucleation rate’s increase rate is greater than that of the growth rate. This is because the generated dislocations cannot be eliminated in time and thus increase, leading to an increase in recrystallization nucleation. Following recrystallization, the grain size is refined. Therefore, increasing the strain amount of the AZ31 Mg alloy facilitates dynamic recrystallization, ultimately resulting in the formation of finer grains.
Figure 2 illustrates the distribution of the velocity and effective strain of Mg alloy sheet processed by ASE. The flow velocity and strain distribution along the thickness direction of AZ31 Mg alloy sheets during the ASE process using the ASE die (L = 4 mm) were analyzed [69]. It can be seen that there is a gradient in both strain and flow velocity along the thickness direction of the sheet. It can be shown that as the Mg alloy enters the shearing deformation zone of the asymmetric extrusion, both strain and flow velocity gradually increase during the extrusion process, with the upper part of the shearing zone having the maximum value. From the simulated results of billet flow velocity in Figure 2b, it can be learned that the effective strains of the upper, middle, and lower parts of the ASE extruded sheets are 4.7, 3.7, and 1.4, respectively, and decrease gradually along the thickness direction of the sheet. The results of the Finite Element Method (FEM) simulation reveal that larger strains occur on the upper surface and smaller dynamic recrystallization grains appear. Moreover, the c-axis orientation of the grains deviates along the ED due to shear deformation.

2.1.2. Differential Speed Extrusion (DSE)

Previous studies have shown that Mg alloy flows smoothly during asymmetric extrusion, and the stress and strain experienced by the metal during the extrusion process are relatively small [55,69]. In order to investigate the microstructure and properties of the AZ31 alloy extruded sheet under greater stress and strain conditions, we designed a differential speed extrusion (DSE).
Figure 3 presents the schematic sectional view and FEM results of the DSE process. The DSE die is designed to induce a substantial flow rate disparity between the upper and lower surfaces of the metal billet, along with a sharp change in flow rate. This generates a larger strain gradient along the thickness direction of the extruded sheets. This can refine the grain size and improve its strength and plasticity. A finite element simulation was conducted to study the velocity and strain distribution along the thickness direction during the DSE process. The results indicate that there is a certain gradient in strain and velocity along the thickness direction of the sheet. The velocity ratio between the upper and lower parts was determined to be 2:1, indicating higher strain and faster velocity at the upper surface. During the extrusion deformation process, the uneven distribution of strain would cause different degrees of dynamic recrystallization in the Mg alloy. The areas with higher strain would undergo greater dynamic recrystallization, resulting in smaller equiaxed grains.
Figure 4 shows the (0002) pole figure and EBSD grain orientation of the CE and DSE sheets. It is evident that the microstructure of the DSE sample is non-uniform along the thickness direction. As shown in Figure 4b, the coarse grains are elongated and deviated from the c-axis of the basal plane, and there are many small dynamically recrystallized (DRX) grains around the elongated grains. Moreover, the basal texture is tilted by about 15° towards the ED. The relationship between the average grain size and the Zener-Hollomon (Z) parameter is expressed as
Ln d = A + B Ln Z
where the temperature-corrected strain rate Z is Z = ε·exp(Q/RT). According to this equation, the larger strain on the upper surface results in a smaller grain size [70,71], with a value of about 8 μm, while the grain size on the lower surface is about 9 μm. Meanwhile, the basal texture on the lower surface is tilted by about 12° towards ED, and the DRX grains on both upper and lower surfaces are tilted in the direction of the applied shear force. Thus, the majority of grains tend to undergo prismatic <a> slip rather than basal <a> slip due to the shear action. This prismatic <a> slip causes the grains to rotate and changes their orientation while increasing the strain between adjacent grains, leading to the generation of secondary stresses between grains, which in turn alters the strain state of each grain. As the Mg alloy undergoes extrusion deformation, when the slip distance of the initial slip system reaches a certain degree, the grain orientation and stress state change significantly, so that the orientation factor of the other slip systems is higher than that of the current system [72,73], thereby altering the activation status of the slip system and ultimately achieving continuous strain.

2.1.3. Normal Gradient Extrusion (NGE)

To further understand the difference in rheological behavior of AZ31 Mg alloy between CE and NGE processes, the stress state of AZ31 Mg alloy during the extrusion process was analyzed as shown in Figure 5. Our research team has previously investigated the microstructure and mechanical properties of Mg alloys prepared by NGE and CE processes [59]. The included angles of the upper and lower dies of the NGE extrusion die are processed at 30°, 45°, 60°, and 90°. In CE symmetrical extrusion, the upper and lower surfaces of the AZ31 alloy sheet in the forming area are subjected to the same force from the die (PT = PB). However, in the NGE non-symmetric extrusion process, the stress on the AZ31 alloy is more complex. When the AZ31 alloy flows into the deformation zone (red zone), it is subjected to a force P applied by the die, which can be divided into two components (PED and PND, respectively). This indicates that the AZ31 alloy bears additional normal stress PND in the NGE extrusion die. This is subjected to different stresses on the upper and lower surfaces of the extruded sheet (PT ≠ PB), resulting in the formation of different flow velocities (VT ≠ VB) on the upper and lower surfaces of the extruded sheet. This is conducive to the formation of additional shear strain along the ED direction during sheet forming [74,75]. Therefore, a large effective strain gradient is formed along the thickness direction of the extruded sheet in the NGE extrusion process. A large effective strain and strain gradient can effectively refine the microstructure of AZ31 alloy sheet and weaken the texture.
Figure 6 illustrates EBSD analysis and (0002) pole figures of the upper surface, middle layer, and lower surface of AZ31 sheet extruded by the NGE-45° process. The texture strength varies in different regions of the same extruded sheet. Specifically, the middle layer of the extruded AZ31 sheet shows a bimodal texture feature, elongated along the ED direction. The texture strength of the middle layer in the NGE-45° sheet is 8.0, reaching its lowest value. Additionally, the basal texture on the upper surface of the GASE-45° sheet is more dispersed and inclined along the ED direction, and new texture components appear along the ED direction. The GASE-45° sheet exhibits lower texture strength in the corresponding region.

2.2. Normal Direction Asymmetric Divergent Die Extrusion Technologies of Mg Alloy

The preparation process of the flat die is relatively simple, but a mismatch exists between the circular cross-sections of the die cavity and the corresponding extrusion cylinder, which is in contrast to the rectangular cross-section of the extruded sheet. As a consequence, uneven deformation occurs along the width direction of the sheet, particularly for sheets with a large aspect ratio. To address this issue and improve the efficiency and quality of extruded sheets, practical production and industrial applications often incorporate flat extrusion cylinders and diverter dies for extruding wide sheets with a large aspect ratio [53,76].

2.2.1. Asymmetric Porthole Die Extrusion (APE)

Figure 7 presents schematic diagrams of the symmetric flow-diverting die and three types of asymmetric flow-diverting dies. Different from conventional dies, the flow-diverting die features an enlarged entry and a flow-diverting baffle at its entrance, allowing for smooth division of the billet into two metal flows during the extrusion process and exposing a new interface. Subsequently, in the high-temperature and high-pressure environment of the die cavity, the newly exposed interfaces can bond tightly to form a good metallurgical bonding interface. Based on the symmetric flow-diverting die, we modified the structure of the flow-diverting baffle to create specific asymmetric flow-diverting dies with different angles (45°, 60°, and 90°). As seen from the geometric shapes of the AZ31 alloy billets, the billet completely fills the die cavity during the extrusion process. The streamline distribution of the alloy shows good symmetry along the ED direction, but there is a difference in streamline angle between the extrusion streamline and the ED direction. Specifically, the streamline angle for symmetrical extrusion is 5°, whereas for the three types of APE asymmetric dies, the streamline angles are 12°, 17°, and 21°, respectively. This indicates that the use of asymmetric flow-diverting dies significantly increases the streamline angle, which gradually increases with the increasing bridge angle of the die. Consequently, the geometric asymmetry of the asymmetric flow-diverting die results in significant asymmetrical flow of the alloy billets [77,78]. Overall, these findings have implications for the design and optimization of flow-diverting dies in the extrusion process.
Figure 8 shows the microstructures and texture evolutions of AZ31 Mg alloy during the CE and APE processes. The CE extruded sheet exhibits typical basal texture characteristics, with the maximum pole density located at the center of the (0002) pole figure and a relatively high maximum pole density value. In contrast, for the three types of asymmetric flow splitting pattern extruded sheets, there are differences not only in the distribution of maximum pole density but also in significant differences in the numerical values of maximum pole density. In terms of the distribution of maximum pole density, all three extruded sheets exhibit a certain degree of angle deviation along the ED direction, with the angle gradually increasing as the flow splitting angle increases, from about 15.3° to about 21.3°. The change in the deviation angle of the maximum pole density along the ED direction is consistent with the change in the asymmetric flow splitting angle. This indicates that introducing an asymmetric flow splitting angle leads to a deviation of the maximum pole density of the extruded sheet along the ED direction. The decrease in the maximum pole density and the more dispersed pole axis indicate that the asymmetric flow splitting pattern extrusion can more effectively weaken the basal texture of magnesium alloy compared with the CE and PE symmetric extrusions, especially for the APE-90 flow splitting pattern extrusion with a large flow splitting angle. To further elucidate the mechanism of the weakening of the texture of APE-90 sheet, the microstructure and texture evolution during CE and APE-90 asymmetric extrusion are analyzed. In the CE extrusion process, every position from 1 to 4 shows uneven microstructure, which is a typical mixed crystal structure. In the APE-90 extrusion process, it gradually transforms from a dynamic recrystallization structure and uneven mixing of undeformed grains in the initial stage of extrusion to a relatively uniform and complete dynamic recrystallization structure, especially at positions 3 and 4, where grain size is smaller, indicating strong shear strain in the final stage of extrusion that promotes dynamic recrystallization. As the extrusion process progresses, the basal slip of the alloy gradually dominates, resulting in the maximum pole density moving from the edge to the center of the pole figure. Compared with the CE extrusion process, the APE-90 extrusion process delays the formation of basal texture and the turning of the pole axis to the central position. APE-90 extruded sheets retain more non-basal orientation grains, forming a weaker basal texture inclined along the extrusion direction, and more dispersed [79,80]. This difference is closely related to the different extrusion channels of the CE symmetric extrusion die and the APE-90 asymmetric extrusion die, which in turn lead to different flow characteristics and shear stresses.

2.2.2. Asymmetric Billet Split Flow Die Extrusion

The development of bimetallic or multicomponent laminated composite materials allows for the integration of the benefits offered by two or more base metal materials. Various bimetallic composite materials, such as AZ31/Al 6061 [81], AZ31/WE43 [82], and AZ31/AZ91 [83], can be prepared through solid-liquid composite. However, high processing temperatures deteriorate their service performance. Multi-layer bimetallic or multicomponent composite materials can be prepared, such as Al/AZ31 [84] and Mg-12Li-1Al/Mg-5Li-1Al (LA121/LA51, wt.%) [85], can be prepared using solid-solid composite methods such as direct co-extrusion [86,87], accumulative roll bonding [85,88], accumulative extrusion-bonding [89], and equal channel angular pressing [90]. Furthermore, appropriate annealing processes can enhance the interfacial binding ability by promoting atomic diffusion between the base metal layers. In our study, the asymmetric deformation extrusion of AZ31 and low rare earth content Mg-0.3 wt.%Y (W0) alloy was achieved by leveraging the differences in material types, building upon the previous symmetric flow splitting die. Figure 9 shows the schematic of the asymmetric billet split (AZ31/W0) flow die extrusion fabrication process. This led to the extrusion preparation of bimetallic laminated composite materials.
Figure 10 shows the microstructure and (0002) pole figure of the longitudinal section of AZ31 sheet, W0 sheet, and AZ31/W0 laminated composite sheet. A notable disparity in microstructure and texture is observed between the AZ31/W0 laminated composite sheet and the single AZ31 sheet and W0 sheet. The average grain sizes of the AZ31 layer and W0 layer in the AZ31/W0 laminated composite sheet are about 18.4 and 9.6 μm, respectively. The composite sheet prepared via symmetric flow splitting die extrusion exhibits a larger average grain size compared with the ordinarily extruded sheet. As shown in Figure 10e–h, the basal texture strength of the AZ31 layer in the AZ31/W0 composite sheet (15.51) is lower and more dispersed compared with the AZ31 sheet with a strong basal texture (23.34). The maximum pole density and distribution of the W0 layer in the AZ31/W0 composite sheet are similar to those of the W0 sheet, with a maximum pole density offset of about ±30° along the ED and a weak orientation distribution along the TD, forming a typical rare earth bimodal texture feature. It can be seen that there is a small interdiffusion zone between the AZ31 layer and the W0 layer, and the width of the interdiffusion zone is about 0.35 μm. According to the selected area electron diffraction (SAED) observation of dark and bright regions, the phases in both regions are magnesium matrix phases, and no compound phase was observed. The high-resolution transmission electron microscopy (HRTEM) image shows a crystallographic interface between the Mg layer and the diffusion zone. The interplanar spacings of {10–10} crystal planes in the matrix and diffusion zone are both 0.160 nm, which can be confirmed as Mg supersaturated solid solution [91,92,93]. The crystal plane angle between the {10–10} crystal plane of the Mg layer and the {10–10} crystal plane of the diffusion zone is about 17°, measured by the crystal plane angle measurement. A small interdiffusion zone is formed between the AZ31 layer and the W0 layer in the AZ31/W0 laminated composite sheet, and there is a good crystallographic matching relationship between the matrix and the diffusion zone. This diffusion zone enables good bonding between the AZ31 layer and the W0 layer.

2.3. Transverse Direction Asymmetric Extrusion Technology for Mg Alloy

Our research team proposed an asymmetric extrusion process along the transverse direction of the sheet. By designing the transverse geometry structure of the extrusion die, we constructed a gradient strain in the transverse direction of the extruded magnesium alloy sheet. Based on optimizing the process parameters, both the basal texture and microstructure of the magnesium alloy extruded sheet were regulated to improve its plastic formability [24,25]. The ultimate goal is to enhance the plasticity of the extruded magnesium alloy sheet through this process.

2.3.1. Asymmetric Extrusion (ASE)

We have designed a transverse gradient asymmetric extrusion flat die, as illustrated in Figure 11, which features an isosceles triangle space at the exit of the die cavity. By using the two sides of the triangle, we are able to regulate the flow velocity difference between the center and edges of the sheet along the transverse direction, resulting in a shear effect and forming asymmetric stress and strain. We conducted various degrees of asymmetric extrusion experiments by adjusting the inclination angle θ (θ = 0°, 15°, 30°, 37°, 45°, 52°, and 60°) of the extrusion die. The die becomes a conventional extrusion (CE) die when θ = 0°, while it becomes a transverse gradient extrusion (TGE) die with different degrees of asymmetry when the inclination angle θ is set to 15°, 30°, 37°, 45°, 52°, and 60°. Figure 11 shows the velocity distribution on the ED-TD plane at the exit of the extrusion die during TGE-52 asymmetric extrusion processes. In the CE process, the velocity direction is mainly aligned with the ED direction. In contrast, in the TGE-52 asymmetric extrusion process, the flow velocity deviates towards TD along ED at the exit of the die, except for the center area of the extruded sheet. This introduces a new flow velocity VTD along TD, which is beneficial in deflecting the basal texture during the extrusion process. Moreover, the angle of deflection from ED to TD increases as VTD increases (VTD = VED × tanθ) and as the inclination angle θ of the die rises.
Figure 12 shows the evolution of microstructure and texture near the extrusion die exit before and after sheets form the TGE Process. During the initial stage of extrusion, the Mg alloy experiences relatively small deformation of its coarse grains, which leads to favorable conditions for the initiation of tensile twinning and results in the formation of many twinned grains [94,95]. At position A, the microstructure is non-uniform, while many small recrystallized grains appear at position B. Additionally, many small green-colored grains with their c-axis inclined along the TD direction can be observed at various locations in the central region of the TGE-52 extruded sheet. The microstructure and texture evolution at the 1/4 edge of the TGE-52 extruded sheet differ significantly from those observed in the central region. The microstructure of the central region of the TGE-52 extruded sheet at the exit of the extrusion die comprises small recrystallized grains. As we move from position F (near the sheet-forming area) to position I (far from the sheet-forming area), the grains become further refined, and a more uniform microstructure is achieved at position I. The texture features of the extruded AZ31Mg alloy sheet exhibit significant variations in different areas. At position F, a basal texture feature with a maximum density of 11.6 is observed. At position G, the basal poles deviate from the ED direction, while at position H, the basal texture component is further reduced. By position I, the basal texture component disappears entirely, and the basal poles deviate greatly from the ED direction, ultimately forming a double peak texture.

2.3.2. Asymmetric Curve Extrusion (ACE)

This work combines the integration of thick and transverse asymmetric structures with the design and fabrication of three-dimensional asymmetric curved extrusion dies, as illustrated in Figure 13. The sheet forming location features upper and lower surfaces designed as circular arcs with different radii of 28 mm and 29 mm, respectively, creating parallel rheological channels of varying lengths. The extrusion velocity direction is deflected from ED to TD at the sheet forming location due to the change of the extrusion die. The AZ31 alloy generates separate velocities along the ED (VED) and TD (VTD) directions during the extrusion process, and the ACE process can effectively introduce additional velocities in the TD direction (VTD). The changes in velocity on the upper and lower surfaces of the sheet are almost identical in the CE process owing to the symmetrical structure of the die. The rheological behavior of AZ31 alloy relative to the middle layer of the extruded sheet shows an asymmetric distribution in the ACE process. The velocity on the upper surface of the extruded sheet is lower than that on the lower surface, and the formation of velocity differences is beneficial for introducing asymmetric shear stress during the extrusion process.
Figure 14 shows the microstructure and texture evolutions of the ACE sheets. ACE alloy billet samples display many small dynamic recrystallized grains that appear in blue and green colors, and the c-axis of these grains is deflected from ND to TD. This asymmetric extrusion process introduces new texture components along the TD. Compared with CE symmetrical extrusion alloy billet samples, the ACE samples show weaker texture strength at the same distance from the sheet forming location. At the same time, the extruded AZ31 alloy sheet obtains a finer microstructure and a weaker basal texture. In addition, the basal pole strength of dynamic recrystallized grains in ACE extruded alloy billet samples (Figure 14a,c–e) is always lower than that in CE alloy billet samples at the same distance from the sheet forming location. This is because the ACE extrusion die has a significant difference in billet flow velocity in the thickness and transverse directions of the sheet, resulting in larger additional shear stress and promoting the dynamic recrystallization process of non-basal-oriented grains, which is manifested as weaker basal texture at the macro level.

3. Conclusions and Outlooks

Conducting research on the strengthening mechanism and formability of Mg alloy sheets holds significant potential for providing superior materials with desirable characteristics such as lightweight, shock absorption, noise reduction, and electromagnetic shielding. In recent years, extensive advancements have been made in high-plastic deformation Mg alloys and plastic processing technology. A novel approach based on asymmetric extrusion technology has been introduced to induce non-symmetric strain in the thickness and transverse directions of the extruded sheet. This method effectively regulates the temperature and stress state during deformation, eliminating the formation of a strong basal texture caused by compression deformation in the sheet’s thickness or transverse direction. Moreover, it allows for the adjustment and control of the texture of Mg alloy sheets, thereby enhancing their subsequent forming properties. Multiple concentrated asymmetric extrusion processing technologies have been developed, including asymmetric extrusion (ASE), differential speed extrusion (DSE), normal gradient extrusion (NGE), transverse gradient extrusion (TGE), asymmetric porthole die extrusion (APE), asymmetric material composition extrusion, and asymmetric curve extrusion (ACE). These innovative extrusion techniques provide a crucial pathway to preparing high-performance Mg alloy sheets, offering a promising strategy for improving the material properties of Mg alloys.
By precisely adjusting the distance and angle between the working band of the mold and the upper and lower surfaces of the sheet, a non-symmetric shear strain gradient has been successfully established along the thickness direction of the sheet (parallel to the surface of the sheet’s thickness). This gradient induces a deviation in the c-axis orientation of the Mg alloy during dynamic recrystallization. Short-process hot extrusion and shear strain effectively weaken the basal texture, significantly improving the mechanical properties of Mg alloy thin sheets. Recent years have witnessed significant progress in the preparation and processing of high-performance Mg alloy sheets.
(1)
Mg-Al-Ca-Mn series microalloyed Mg alloys have been developed, whereby adding rare earth elements such as Ce, Y, and Gd, even at low concentrations, strongly weakens the basal textures.
(2)
Plastic processing technologies such as equal channel angular rolling (ECAR) and pre-deformation control have been developed, introducing gradient strain into the sheet plane, which is conducive to a large amount of basal slip and tensile twinning opening. As a result, forming a c-axis//RD texture orientation feature with a certain {10-12} twin structure significantly enhances the room temperature formability of Mg alloy sheets.
(3)
Wide-width Mg sheet near-isothermal rolling technology has been developed, realizing high-precision rolling of large coil weight wide-width Mg alloy sheets rolls and significantly improving the rolling formability, organization, and performance uniformity of Mg alloy sheets.
In order to achieve efficient preparation of high-formability Mg alloy sheets with weak basal texture and low isotropy, future work should focus on the following aspects.
(1)
The development of low-cost, low-content Mg-Al series Mg alloys and their sheet processing and preparation technology is crucial. This can be achieved by regulating crystal orientation through alloy elements to improve the balance between Mg alloy strength and formability.
(2)
Optimizing the plastic deformation strain path and prefabricating the twinning orientation of Mg alloy sheets is necessary. Coupling with Mg alloy recrystallization behavior can form crystal orientations favorable for Mg alloy plastic deformation, ultimately controlling the isotropy and formability of Mg alloy sheets.
(3)
Exploring the activity of non-basal <a> dislocations and <c+a> dislocations through plastic deformation strain, further systematically theorizing and experimentally verifying, quantitatively analyzing the relationship between dislocation activity and plastic deformation mechanism, and predicting the formability of Mg alloys.
(4)
The development of advanced deformation Mg alloy extrusion die design and comprehensive processing technology, focusing on high strength and toughness, is of paramount importance. It is crucial to establish efficient production and processing technologies for wide-width Mg alloy profiles, along with the refinement of high-precision profile heat treatment, straightening, and other finishing techniques and equipment. Furthermore, the exploration of ultra-wide and high-precision deformation Mg alloy profiles should also be a significant area of investigation.

Author Contributions

Conceptualization, Q.Y., B.J. and F.P.; methodology, D.Z.; investigation, P.P. and G.W.; writing—original draft preparation, Q.Y.; writing—review and editing, J.Z. and B.J.; supervision, F.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China (52271091, 52271092), Chongqing Science and Technology Commission (CSTB2022 NSCQ-MSX0891).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

There is no conflict of interest with any person or foundation.

References

  1. Zemkova, M.; Minarik, P.; Dittrich, J.; Bohlen, J.; Kral, R. Individual effect of Y and Nd on the microstructure formation of Mg-Y-Nd alloys processed by severe plastic deformation and their effect on the subsequent mechanical and corrosion properties. J. Magn. Alloys 2023, 11, 509–521. [Google Scholar] [CrossRef]
  2. Wang, S.; Pan, J.; Xie, W.; Yang, J.; Zhang, W.; Chen, W. Effects of Extrusion Ratio on the Microstructure, Texture and Mechanical Properties of Mg-2.5Nd-0.5Zn-0.5Zr Alloy Sheets. J. Mater. Eng. Perform. 2023, 32, 4834–4845. [Google Scholar] [CrossRef]
  3. Zhang, B.; Wang, Y.; Geng, L.; Lu, C. Effects of calcium on texture and mechanical properties of hot-extruded Mg-Zn-Ca alloys. Mater. Sci. Eng. A 2012, 539, 56–60. [Google Scholar] [CrossRef]
  4. Yang, Q.; Jiang, B.; Song, B.; Yu, Z.; He, D.; Chai, Y.; Zhang, J.; Pan, F. The effects of orientation control via tension-compression on microstructural evolution and mechanical behavior of AZ31 Mg alloy sheet. J. Magn. Alloys 2022, 10, 411–422. [Google Scholar] [CrossRef]
  5. Bian, M.Z.; Sasaki, T.T.; Suh, B.C.; Nakata, T.; Kamado, S.; Hono, K. A heat-treatable Mg-Al-Ca-Mn-Zn sheet alloy with good room temperature formability. Scr. Mater. 2017, 138, 151–155. [Google Scholar] [CrossRef]
  6. Jin, H.; Amirkhiz, B.S.; Lloyd, D.J. Improvement of Superplasticity in High-Mg Aluminum Alloys by Sacrifice of Some Room Temperature Formability. Metall. Mater. Trans. A 2018, 49, 1962–1979. [Google Scholar] [CrossRef]
  7. Yu, H.; Li, C.; Xin, Y.; Chapuis, A.; Huang, X.; Liu, Q. The mechanism for the high dependence of the Hall-Petch slope for twinning/slip on texture in Mg alloys. Acta Mater. 2017, 128, 313–326. [Google Scholar] [CrossRef]
  8. Sabokpa, O.; Zarei-Hanzaki, A.; Abedi, H.R. An investigation into the hot ductility behavior of AZ81 magnesium alloy. Mater. Sci. Eng. A 2012, 550, 31–38. [Google Scholar] [CrossRef]
  9. Wu, Z.; Ahmad, R.; Yin, B.; Sandlöbes, S.; Curtin, W.A. Mechanistic origin and prediction of enhanced ductility in magnesium alloys. Science 2018, 359, 447–452. [Google Scholar] [CrossRef] [Green Version]
  10. Basu, S.; Dogan, E.; Kondori, B.; Karaman, I.; Benzerga, A.A. Towards designing anisotropy for ductility enhancement: A theory-driven investigation in Mg-alloys. Acta Mater. 2017, 131, 349–362. [Google Scholar] [CrossRef] [Green Version]
  11. Yi, S.B.; Davies, C.H.J.; Brokmeier, H.G.; Bolmaro, R.E.; Kainer, K.U.; Homeyer, J. Deformation and texture evolution in AZ31 magnesium alloy during uniaxial loading. Acta Mater. 2006, 54, 549–562. [Google Scholar] [CrossRef]
  12. Agnew, S.R.; Duygulu, Ö. Plastic anisotropy and the role of non-basal slip in magnesium alloy AZ31B. Int. J. Plast. 2005, 21, 1161–1193. [Google Scholar] [CrossRef]
  13. Agnew, S.R.; Singh, A.; Calhoun, C.A.; Mulay, R.P.; Bhattacharyya, J.J.; Somekawa, H.; Mukai, T.; Clausen, B.; Wu, P.D. In-situ neutron diffraction of a quasicrystal-containing Mg alloy interpreted using a new polycrystal plasticity model of hardening due to {10.2} tensile twinning. Int. J. Plast. 2018, 100, 34–51. [Google Scholar] [CrossRef]
  14. Zhang, D.; Zhang, D.; Xu, T.; Chen, S.; Zhang, Y.; Li, X.; Zhang, J. Achieving high-strength in Mg-0.8Zn-0.2Zr (wt.%) alloy extruded at low temperature. Mater. Sci. Eng. A-Struct. Mater. Prop. Microstruct. Process. 2021, 822, 141657. [Google Scholar] [CrossRef]
  15. Yang, Q.; Jiang, B.; Song, B.; Yu, D.; Chai, S.; Zhang, J.; Pan, F. Mechanical behavior and microstructure evolution for extruded AZ31 sheet under side direction strain. Prog. Nat. Sci.-Mater. Int. 2020, 30, 270–277. [Google Scholar] [CrossRef]
  16. Zhou, C.; Liang, G.; Liu, Y.; Zhang, H.; Chen, X.; Li, Y. Pore structure of porous Mg-1Mn-xZn alloy fabricated by metal-gas eutectic unidirectional solidification. J. Magn. Alloys 2022, 10, 2137–2146. [Google Scholar] [CrossRef]
  17. Zhu, S.Q.; Ringer, S.P. On the role of twinning and stacking faults on the crystal plasticity and grain refinement in magnesium alloys. Acta Mater. 2018, 144, 365–375. [Google Scholar] [CrossRef]
  18. Zhang, X.; Kevorkov, D.; Jung, I.-H.; Pekguleryuz, M. Phase equilibria on the ternary Mg-Mn-Ce system at the Mg-rich corner. J. Alloys Compd. 2009, 482, 420–428. [Google Scholar] [CrossRef]
  19. Ji, H.; Wu, G.; Liu, W.; Sun, J.; Ding, W. Role of extrusion temperature on the microstructure evolution and tensile properties of an ultralight Mg-Li-Zn-Er alloy. J. Alloys Compd. 2021, 876, 160181. [Google Scholar] [CrossRef]
  20. Li, R.G.; Li, H.R.; Pan, H.C.; Xie, D.S.; Zhang, J.H.; Fang, D.Q.; Dai, Y.Q.; Zhao, D.Y.; Zhang, H. Achieving exceptionally high strength in binary Mg-13Gd alloy by strong texture and substantial precipitates. Scr. Mater. 2021, 193, 142–146. [Google Scholar] [CrossRef]
  21. Li, Y.; Nie, K.; Deng, K.; Yang, A. Microstructures and Mechanical Properties of Low-Alloyed Mg-Zn-Y Magnesium Alloy. Rare Metall. Mater. Eng. 2021, 50, 1425–1432. [Google Scholar]
  22. Wang, S.; Zhang, W.; Yang, J.; Pan, J.; Wang, H.; Chen, W.; Cui, G. Evolution of Microstructures, Texture, Damping and Mechanical Properties of Hot Extruded Mg-Nd-Zn-Zr Alloy. J. Mater. Eng. Perform. 2021, 30, 8872–8882. [Google Scholar] [CrossRef]
  23. Zhang, C.; Peng, C.; Huang, J.; Zhao, Y.; Han, T.; Wang, G.; Wu, L.; Huang, G. Improving Mechanical Properties of Mg-Sc Alloy by Surface AZ31 Layer. Metals 2021, 11, 2021. [Google Scholar] [CrossRef]
  24. Zhuang, Y.; Zhang, Y.; Zeng, Q.; Li, J. Coupling the semi-solid treatment and hot extrusion to strengthen a Mg-Zn-Gd alloy containing I-phase. Mater. Lett. 2021, 287, 129294. [Google Scholar] [CrossRef]
  25. Bairagi, D.; Mandal, S. A comprehensive review on biocompatible Mg-based alloys as temporary orthopaedic implants: Current status, challenges, and future prospects. J. Magn. Alloys 2022, 10, 627–669. [Google Scholar] [CrossRef]
  26. Suzuki, A.; Saddock, N.; Jones, J.; Pollock, T. Structure and transition of eutectic (Mg,Al)Ca Laves phase in a die-cast Mg-Al-Ca base alloy. Scr. Mater. 2004, 51, 1005–1010. [Google Scholar] [CrossRef]
  27. Che, B.; Lu, L.; Kang, W.; Zhong, Y.; Ma, M.; Liu, L.; Wu, Z. Effect of Expansion Sphere Diameter on Deformation Behavior of AZ31 Mg Alloy during Extrusion. J. Mater. Eng. Perform. 2022, 31, 8512–8521. [Google Scholar] [CrossRef]
  28. Du, P.; Mei, D.; Furushima, T.; Zhu, S.; Wang, L.; Zhou, Y.; Guan, S. In vitro corrosion properties of HTHEed Mg-Zn-Y-Nd alloy microtubes for stent applications: Influence of second phase particles and crystal orientation. J. Magn. Alloys 2022, 10, 1286–1295. [Google Scholar] [CrossRef]
  29. Zhang, D.; Pan, H.; Li, J.; Xie, D.; Zhang, D.; Che, C.; Meng, J.; Qin, G. Fabrication of exceptionally high-strength Mg-4Sm-0.6Zn-0.4Zr alloy via low-temperature extrusion. Mater. Sci. Eng. A-Struct. Mater. Prop. Microstruct. Process. 2022, 833, 142565. [Google Scholar] [CrossRef]
  30. Xu, Y.; Li, J.; Qi, M.; Guo, W.; Deng, Y. A newly developed Mg-Zn-Gd-Mn-Sr alloy for degradable implant applications: Influence of extrusion temperature on microstructure, mechanical properties and in vitro corrosion behavior. Mater. Charact. 2022, 188, 111867. [Google Scholar] [CrossRef]
  31. Zhai, Y.; Hou, X.; Yuan, Z.; Zhang, P.; Guan, Q. Analysis of Crystallographic Texture and Mechanical Anisotropy of an Extruded Mg-RE Alloy. Rare Metall. Mater. Eng. 2018, 47, 1341–1346. [Google Scholar] [CrossRef]
  32. Meng, S.J.; Yu, H.; Fan, S.D.; Kim, Y.M.; Park, S.H.; Zhao, W.M.; You, B.S.; Shin, K.S. A high-ductility extruded Mg-Bi-Ca alloy. Mater. Lett. 2020, 261, 127066. [Google Scholar] [CrossRef]
  33. Bian, M.; Huang, X.; Mabuchi, M.; Chino, Y. Compositional optimization of Mg-Zn-Sc sheet alloys for enhanced room temperature stretch formability. J. Alloys Compd. 2020, 818, 152891. [Google Scholar] [CrossRef]
  34. Jiang, M.G.; Xu, C.; Nakata, T.; Yan, H.; Chen, R.S.; Kamado, S. Enhancing strength and ductility of Mg-Zn-Gd alloy via slow-speed extrusion combined with pre-forging. J. Alloys Compd. 2017, 694, 1214–1223. [Google Scholar] [CrossRef] [Green Version]
  35. Wang, H.Y.; Yu, Z.P.; Zhang, L.; Liu, C.G.; Zha, M.; Wang, C.; Jiang, Q.C. Achieving high strength and high ductility in magnesium alloy using hard-plate rolling (HPR) process. Sci. Rep. 2015, 5, 17100. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Kim, W.J.; Hong, S.I.; Lee, J.M.; Kim, S.H. Dispersion of TiC particles in an in situ aluminum matrix composite by shear plastic flow during high-ratio differential speed rolling. Mater. Sci. Eng. A 2013, 559, 325–332. [Google Scholar] [CrossRef]
  37. Chang, L.L.; Kang, S.B.; Cho, J.H. Influence of strain path on the microstructure evolution and mechanical properties in AM31 magnesium alloy sheets processed by differential speed rolling. Mater. Des. 2013, 44, 144–148. [Google Scholar] [CrossRef]
  38. Naghdi, F.; Mahmudi, R.; Kang, J.Y.; Kim, H.S. Contributions of different strengthening mechanisms to the shear strength of an extruded Mg-4Zn-0.5Ca alloy. Philos. Mag. 2015, 95, 3452–3466. [Google Scholar] [CrossRef]
  39. Mao, B.; Li, B.; Lin, D.; Liao, Y. Enhanced room temperature stretch formability of AZ31B magnesium alloy sheet by laser shock peening. Mater. Sci. Eng. A 2019, 756, 219–225. [Google Scholar] [CrossRef]
  40. Lee, J.U.; Kim, S.-H.; Kim, Y.J.; Park, S.H. Improvement in bending formability of rolled magnesium alloy through precompression and subsequent annealing. J. Alloys Compd. 2019, 787, 519–526. [Google Scholar] [CrossRef]
  41. Apachitei, I.; Fratila-Apachitei, L.E.; Duszczyk, J. Microgalvanic activity of an Mg-Al-Ca-based alloy studied by scanning Kelvin probe force microscopy. Scr. Mater. 2007, 57, 1012–1015. [Google Scholar] [CrossRef]
  42. Barnett, M.R. Importance of propagation in controlling the twinning stress in Mg. Scr. Mater. 2019, 162, 447–450. [Google Scholar] [CrossRef]
  43. Liu, X.-Y.; Lu, L.-W.; Sheng, K.; Zhou, T. Microstructure and Texture Evolution during the Direct Extrusion and Bending–Shear Deformation of AZ31 Magnesium Alloy. Acta Metall. Sin.-Engl. Lett. 2018, 32, 710–718. [Google Scholar] [CrossRef] [Green Version]
  44. Beygelzimer, Y.; Kulagin, R.; Estrin, Y.; Toth, L.S.; Kim, H.S.; Latypov, M.I. Twist Extrusion as a Potent Tool for Obtaining Advanced Engineering Materials: A Review. Adv. Eng. Mater. 2017, 19, 1600873. [Google Scholar] [CrossRef]
  45. Wang, F.; Zheng, R.; Chen, J.; Lyu, S.; Li, Y.; Xiao, W.; Ma, C. Significant improvement in the strength of Mg-Al-Zn-Ca-Mn extruded alloy by tailoring the initial microstructure. Vacuum 2019, 161, 429–433. [Google Scholar] [CrossRef]
  46. Zheng, L.; Nie, H.; Zhang, W.; Liang, W.; Wang, Y. Microstructural refinement and improvement of mechanical properties of hot-rolled Mg-3Al-Zn alloy sheets subjected to pre-extrusion and Al-Si alloying. Mater. Sci. Eng. A 2018, 722, 58–68. [Google Scholar] [CrossRef]
  47. Zhao, L.; Xin, Y.; Wu, Y.; Liu, Q. The texture dependence of strength in slip and twinning predominant deformations of Mg-3Al-1Zn alloy. Mater. Sci. Eng. A 2018, 717, 34–40. [Google Scholar] [CrossRef]
  48. Yang, H.W.; Widiantara, I.P.; Ko, Y.G. Effect of deformation path on texture and tension properties of submicrocrystalline Al-Mg-Si alloy fabricated by differential speed rolling. Mater. Lett. 2018, 213, 54–57. [Google Scholar] [CrossRef]
  49. Jin, S.-C.; Cha, J.W.; Lee, J.H.; Lee, T.; Han, S.H.; Park, S.H. Improvement in tensile strength of extruded Mg-5Bi alloy through addition of Sn and its underlying strengthening mechanisms. J. Magn. Alloys 2022, 10, 3100–3112. [Google Scholar] [CrossRef]
  50. Jeong, H.T.; Kim, W.J. Critical review of superplastic magnesium alloys with emphasis on tensile elongation behavior and deformation mechanisms. J. Magn. Alloys 2022, 10, 1133–1153. [Google Scholar] [CrossRef]
  51. Jafari, H.; Tehrani, A.H.M.; Tehrani, M.; Heydari, M. Effect of extrusion process on microstructure and mechanical and corrosion properties of biodegradable Mg-5Zn-1.5Y magnesium alloy. Int. J. Miner. Metall. Mater. 2022, 29, 490–502. [Google Scholar] [CrossRef]
  52. Zhou, M.; Huang, X.; Morisada, Y.; Fujii, H.; Chino, Y. Effects of Ca and Sr additions on microstructure, mechanical properties, and ignition temperature of hot-rolled Mg-Zn alloy. Mater. Sci. Eng. A 2020, 769, 138474. [Google Scholar] [CrossRef]
  53. Zhou, X.; Ha, C.; Yi, S.; Bohlen, J.; Schell, N.; Chi, Y.; Zheng, M.; Brokmeier, H.-G. Texture and Lattice Strain Evolution during Tensile Loading of Mg-Zn Alloys Measured by Synchrotron Diffraction. Metals 2020, 10, 124. [Google Scholar] [CrossRef] [Green Version]
  54. Zhang, J.; Liu, S.; Wu, R.; Hou, L.; Zhang, M. Recent developments in high-strength Mg-RE-based alloys: Focusing on Mg-Gd and Mg-Y systems. J. Magn. Alloys 2018, 6, 277–291. [Google Scholar] [CrossRef]
  55. Yang, Q.; Jiang, B.; Tian, Y.; Liu, W.; Pan, F. A tilted weak texture processed by an asymmetric extrusion for magnesium alloy sheets. Mater. Lett. 2013, 100, 29–31. [Google Scholar] [CrossRef]
  56. Yang, Q.; Jiang, B.; Zhou, G.; Dai, J.; Pan, F. Influence of an asymmetric shear deformation on microstructure evolution and mechanical behavior of AZ31 magnesium alloy sheet. Mater. Sci. Eng. A 2014, 590, 440–447. [Google Scholar] [CrossRef]
  57. Yang, Q.-S.; Jiang, B.; Yu, Z.-J.; Dai, Q.-W.; Luo, S.-Q. Effect of Extrusion Strain Path on Microstructure and Properties of AZ31 Magnesium Alloy Sheet. Acta Metall. Sin.-Engl. Lett. 2015, 28, 1257–1263. [Google Scholar] [CrossRef]
  58. Yang, Q.; Jiang, B.; He, J.; Song, B.; Liu, W.; Dong, H.; Pan, F. Tailoring texture and refining grain of magnesium alloy by differential speed extrusion process. Mater. Sci. Eng. A 2014, 612, 187–191. [Google Scholar] [CrossRef]
  59. Xu, J.; Yang, T.; Jiang, B.; Song, J.; He, J.; Wang, Q.; Chai, Y.; Huang, G.; Pan, F. Improved mechanical properties of Mg-3Al-1Zn alloy sheets by optimizing the extrusion die angles: Microstructural and texture evolution. J. Alloys Compd. 2018, 762, 719–729. [Google Scholar] [CrossRef]
  60. Xu, J.; Liu, W.; Jiang, B.; Yang, H.; Li, X.; Kang, Y.; Zhou, N.; Zhang, W.; Zheng, K.; Pan, F. Forming novel texture and enhancing the formability in Mg-3Al-Zn alloy sheets fabricated by transverse gradient extrusion. J. Mater. Res. Technol. 2022, 18, 3143–3149. [Google Scholar] [CrossRef]
  61. Wang, Q.; Song, J.; Jiang, B.; Tang, A.; Chai, Y.; Yang, T.; Huang, G.; Pan, F. An investigation on microstructure, texture and formability of AZ31 sheet processed by asymmetric porthole die extrusion. Mater. Sci. Eng. A 2018, 720, 85–97. [Google Scholar] [CrossRef]
  62. Wang, Q.; Shen, Y.; Jiang, B.; Tang, A.; Song, J.; Jiang, Z.; Yang, T.; Huang, G.; Pan, F. Enhanced stretch formability at room temperature for Mg-Al-Zn/Mg-Y laminated composite via porthole die extrusion. Mater. Sci. Eng. A 2018, 731, 184–194. [Google Scholar] [CrossRef]
  63. Xu, J.; Jiang, B.; Kang, Y.; Zhao, J.; Zhang, W.; Zheng, K.; Pan, F. Tailoring microstructure and texture of Mg-3Al-1Zn alloy sheets through curve extrusion process for achieving low planar anisotropy. J. Mater. Sci. Technol. 2022, 113, 48–60. [Google Scholar] [CrossRef]
  64. Wang, Z.; Gu, R.; Chen, S.; Wang, W.; Wei, X. Effect of upper-die temperature on the formability of AZ31B magnesium alloy sheet in stamping. J. Mater. Process. Technol. 2018, 257, 180–190. [Google Scholar] [CrossRef]
  65. Wang, W.; Ma, L.; Chai, S.; Zhang, W.; Chen, W.; Feng, Y.; Cui, G. Role of one direction strong texture in stretch formability for ZK60 magnesium alloy sheet. Mater. Sci. Eng. A 2018, 730, 162–167. [Google Scholar] [CrossRef]
  66. Sun, J.; Yang, Z.; Liu, H.; Han, J.; Wu, Y.; Zhuo, X.; Song, D.; Jiang, J.; Ma, A.; Wu, G. Tension-compression asymmetry of the AZ91 magnesium alloy with multi-heterogenous microstructure. Mater. Sci. Eng. A 2019, 759, 703–707. [Google Scholar] [CrossRef]
  67. Xu, J.; Song, J.; Jiang, B.; He, J.; Wang, Q.; Liu, B.; Huang, G.; Pan, F. Effect of effective strain gradient on texture and mechanical properties of Mg-3Al-1Zn alloy sheets produced by asymmetric extrusion. Mater. Sci. Eng. A-Struct. Mater. Prop. Microstruct. Process. 2017, 706, 172–180. [Google Scholar] [CrossRef]
  68. Jin, Z.-Z.; Zha, M.; Wang, S.-Q.; Wang, S.-C.; Wang, C.; Jia, H.-L.; Wang, H.-Y. Alloying design and microstructural control strategies towards developing Mg alloys with enhanced ductility. J. Magn. Alloys 2022, 10, 1191–1206. [Google Scholar] [CrossRef]
  69. Yang, Q.S.; Jiang, B.; Dai, J.H.; Xiang, Q.; Pan, F.S. Microstructure and mechanical behaviour of asymmetric extruded Mg-3Al-1Zn alloy sheets. Mater. Sci. Technol. 2013, 29, 710–714. [Google Scholar] [CrossRef]
  70. Kim, W.J.; Yoo, S.J.; Chen, Z.H.; Jeong, H.T. Grain size and texture control of Mg-3Al-1Zn alloy sheet using a combination of equal-channel angular rolling and high-speed-ratio differential speed-rolling processes. Scr. Mater. 2009, 60, 897–900. [Google Scholar] [CrossRef]
  71. Huang, X.; Suzuki, K.; Watazu, A.; Shigematsu, I.; Saito, N. Microstructural and textural evolution of AZ31 magnesium alloy during differential speed rolling. J. Alloys Compd. 2009, 479, 726–731. [Google Scholar] [CrossRef]
  72. Stanford, N.; Sha, G.; Xia, J.H.; Ringer, S.P.; Barnett, M.R. Solute segregation and texture modification in an extruded magnesium alloy containing gadolinium. Scr. Mater. 2011, 65, 919–921. [Google Scholar] [CrossRef]
  73. Stanford, N.; Taylor, A.S.; Cizek, P.; Siska, F.; Ramajayam, M.; Barnett, M.R. Twinning in magnesium-based lamellar microstructures. Scr. Mater. 2012, 67, 704–707. [Google Scholar] [CrossRef]
  74. Suzuki, A.; Saddock, N.D.; Jones, J.W.; Pollock, T.M. Solidification paths and eutectic intermetallic phases in Mg-Al-Ca ternary alloys. Acta Mater. 2005, 53, 2823–2834. [Google Scholar] [CrossRef]
  75. Tekumalla, S.; Bibhanshu, N.; Suwas, S.; Gupta, M. Superior ductility in magnesium alloy-based nanocomposites: The crucial role of texture induced by nanoparticles. J. Mater. Sci. 2019, 54, 8711–8718. [Google Scholar] [CrossRef]
  76. Pan, H.; Wang, F.; Feng, M.; Jin, L.; Dong, J.; Wu, P. Mechanical behavior and microstructural evolution in rolled Mg-3Al-1Zn-0.5Mn alloy under large strain simple shear. Mater. Sci. Eng. A 2018, 712, 585–591. [Google Scholar] [CrossRef]
  77. Guan, D.; Rainforth, W.M.; Gao, J.; Sharp, J.; Wynne, B.; Ma, L. Individual effect of recrystallisation nucleation sites on texture weakening in a magnesium alloy: Part 1—Double twins. Acta Mater. 2017, 135, 14–24. [Google Scholar] [CrossRef]
  78. Cepeda-Jiménez, C.M.; Prado-Martínez, C.; Pérez-Prado, M.T. Understanding the high temperature reversed yield asymmetry in a Mg-rare earth alloy by slip trace analysis. Acta Mater. 2018, 145, 264–277. [Google Scholar] [CrossRef]
  79. Shi, H.; Xu, C.; Hu, X.; Gan, W.; Wu, K.; Wang, X. Improving the Young’s modulus of Mg via alloying and compositing—A short review. J. Magn. Alloys 2022, 10, 2009–2024. [Google Scholar] [CrossRef]
  80. Pulido-Gonzalez, N.; Hidalgo-Manrique, P.; Garcia-Rodriguez, S.; Torres, B.; Rams, J. Effect of heat treatment on the mechanical and biocorrosion behaviour of two Mg-Zn-Ca alloys. J. Magn. Alloys 2022, 10, 540–554. [Google Scholar] [CrossRef]
  81. Liu, J.C.; Hu, J.; Nie, X.Y.; Li, H.X.; Du, Q.; Zhang, J.S.; Zhuang, L.Z. The interface bonding mechanism and related mechanical properties of Mg/Al compound materials fabricated by insert molding. Mater. Sci. Eng. A-Struct. Mater. Prop. Microstruct. Process. 2015, 635, 70–76. [Google Scholar] [CrossRef]
  82. Zhao, K.N.; Li, H.X.; Luo, J.R.; Liu, Y.J.; Du, Q.; Zhang, J.S. Interfacial bonding mechanism and mechanical properties of novel AZ31/WE43 bimetal composites fabricated by insert molding method. J. Alloys Compd. 2017, 729, 344–353. [Google Scholar] [CrossRef]
  83. Zhao, K.N.; Liu, J.C.; Nie, X.Y.; Li, Y.; Li, H.X.; Du, Q.; Zhuang, L.Z.; Zhang, J.S. Interface formation in magnesium-magnesium bimetal composites fabricated by insert molding method. Mater. Des. 2016, 91, 122–131. [Google Scholar] [CrossRef]
  84. Wu, D.; Chen, R.-S.; Han, E.-H. Bonding interface zone of Mg-Gd-Y/Mg-Zn-Gd laminated composite fabricated by equal channel angular extrusion. Trans. Nonferrous Met. Soc. China 2010, 20, S613–S618. [Google Scholar] [CrossRef]
  85. Thirumurugan, M.; Rao, S.A.; Kumaran, S.; Rao, T.S. Improved ductility in ZM21 magnesium-aluminium macrocomposite produced by co-extrusion. J. Mater. Process. Technol. 2011, 211, 1637–1642. [Google Scholar] [CrossRef]
  86. Liu, X.B.; Chen, R.S.; Han, E.H. Preliminary investigations on the Mg-Al-Zn/Al laminated composite fabricated by equal channel angular extrusion. J. Mater. Process. Technol. 2009, 209, 4675–4681. [Google Scholar] [CrossRef]
  87. Wu, H.; Wang, T.; Wu, R.; Hou, L.; Zhang, J.; Li, X.; Zhang, M. Effects of Annealing Process on the Interface of Alternate alpha/beta Mg-Li Composite Sheets Prepared by Accumulative Roll Bonding. J. Mater. Process. Technol. 2018, 254, 265–276. [Google Scholar] [CrossRef]
  88. Negendank, M.; Mueller, S.; Reimers, W. Coextrusion of Mg-Al macro composites. J. Mater. Process. Technol. 2012, 212, 1954–1962. [Google Scholar] [CrossRef]
  89. Mozaffari, A.; Manesh, H.D.; Janghorban, K. Evaluation of mechanical properties and structure of multilayered Al/Ni composites produced by accumulative roll bonding (ARB) process. J. Alloys Compd. 2010, 489, 103–109. [Google Scholar] [CrossRef]
  90. Xin, Y.; Hong, R.; Feng, B.; Yu, H.; Wu, Y.; Liu, Q. Fabrication of Mg/AL multilayer plates using an accumulative extrusion bonding process. Mater. Sci. Eng. A-Struct. Mater. Prop. Microstruct. Process. 2015, 640, 210–216. [Google Scholar] [CrossRef]
  91. Meng, F.; Lv, S.; Yang, Q.; Qiu, X.; Yan, Z.; Duan, Q.; Meng, J. Multiplex intermetallic phases in a gravity die-cast Mg-6.0Zn-1.5Nd-0.5Zr (wt%) alloy. J. Magn. Alloys 2022, 10, 209–223. [Google Scholar] [CrossRef]
  92. Luginin, N.A.; Eroshenko, A.Y.; Legostaeva, E.V.; Schmidt, J.; Tolmachev, A.I.; Uvarkin, P.V.; Sharkeev, Y.P. Effect of Severe Plastic Deformation by Extrusion on Microstructure and Physical and Mechanical Properties of Mg-Y-Nd and Mg-Ca Alloys. Technol. Phys. 2022, 67, 791–797. [Google Scholar] [CrossRef]
  93. Ling, L.; Cai, S.; Li, Q.; Sun, J.; Bao, X.; Xu, G. Recent advances in hydrothermal modification of calcium phosphorus coating on magnesium alloy. J. Magn. Alloys 2022, 10, 62–80. [Google Scholar] [CrossRef]
  94. Guo, L.; Wang, J.; Yun, X.; Chen, Z. Fabrication of aluminum-magnesium clad composites by continuous extrusion. Mater. Sci. Eng. A-Struct. Mater. Prop. Microstruct. Process. 2021, 802, 140670. [Google Scholar] [CrossRef]
  95. Bao, J.; Li, Q.; Chen, X.; Zhang, Q.; Chen, Z. Microstructure and texture evolution with Sm addition in extruded Mg-Gd-Sm-Zr alloy. Mater. Res. Express 2021, 8, 096523. [Google Scholar] [CrossRef]
Figure 1. (a) Schematic section view, (b) EBSD orientation maps and (c) (0002) pole figures of conventional extrusion and asymmetric extrusion [55].
Figure 1. (a) Schematic section view, (b) EBSD orientation maps and (c) (0002) pole figures of conventional extrusion and asymmetric extrusion [55].
Materials 16 05255 g001
Figure 2. The distribution of the velocity (a) and effective strain (b) of Mg alloy sheet processed by ASE [55].
Figure 2. The distribution of the velocity (a) and effective strain (b) of Mg alloy sheet processed by ASE [55].
Materials 16 05255 g002
Figure 3. Schematic sectional view (a), FEM results of the extrusion process (b), the velocity (c) and effective strain (d) [58].
Figure 3. Schematic sectional view (a), FEM results of the extrusion process (b), the velocity (c) and effective strain (d) [58].
Materials 16 05255 g003
Figure 4. (0002) Pole figures and EBSD orientation maps of CE sample (a), the DSE sample at top surface (b), mid-layer (c) and bottom surface (d) [58].
Figure 4. (0002) Pole figures and EBSD orientation maps of CE sample (a), the DSE sample at top surface (b), mid-layer (c) and bottom surface (d) [58].
Materials 16 05255 g004
Figure 5. Schematic sectional view of CE (a) and NGE (b), (c) the analysis of AZ31 during NGE processes [59].
Figure 5. Schematic sectional view of CE (a) and NGE (b), (c) the analysis of AZ31 during NGE processes [59].
Materials 16 05255 g005
Figure 6. EBSD and (0002) pole figures of NGE-45° sheet: (a,d) upper surface, (b,e) middle layer, (c,f) lower surface [59].
Figure 6. EBSD and (0002) pole figures of NGE-45° sheet: (a,d) upper surface, (b,e) middle layer, (c,f) lower surface [59].
Materials 16 05255 g006
Figure 7. Schematic drawing of longitudinal section of extrusion die: (a) symmetric porthole die; (b) asymmetric 45° porthole die; (c) asymmetric 60° porthole die; (d) asymmetric 90° porthole die [61].
Figure 7. Schematic drawing of longitudinal section of extrusion die: (a) symmetric porthole die; (b) asymmetric 45° porthole die; (c) asymmetric 60° porthole die; (d) asymmetric 90° porthole die [61].
Materials 16 05255 g007
Figure 8. The microstructures and texture evolutions of AZ31 Mg alloy during the CE and APE-90 processes [61].
Figure 8. The microstructures and texture evolutions of AZ31 Mg alloy during the CE and APE-90 processes [61].
Materials 16 05255 g008
Figure 9. Schematic of the asymmetric billet split (AZ31/W0) flow die extrusion fabrication process [62].
Figure 9. Schematic of the asymmetric billet split (AZ31/W0) flow die extrusion fabrication process [62].
Materials 16 05255 g009
Figure 10. Microstructure and microscopic texture ((0002)) of longitudinal section of AZ31-W0 (1), AZ31/W0 sheets (2). (a) TEM image of the AZ31/W0 interface at low magnification; (b) TEM image of interface between diffusion zone and Mg layer; (c) HRTEM image of yellow frame A in (b); (d) the high magnification view of HRTEM image of yellow frame B in (c) [62].
Figure 10. Microstructure and microscopic texture ((0002)) of longitudinal section of AZ31-W0 (1), AZ31/W0 sheets (2). (a) TEM image of the AZ31/W0 interface at low magnification; (b) TEM image of interface between diffusion zone and Mg layer; (c) HRTEM image of yellow frame A in (b); (d) the high magnification view of HRTEM image of yellow frame B in (c) [62].
Materials 16 05255 g010
Figure 11. (a) The schematic section of TGE die; (b) Flow velocity distribution near the die exit of the AZ31 Mg alloy [60].
Figure 11. (a) The schematic section of TGE die; (b) Flow velocity distribution near the die exit of the AZ31 Mg alloy [60].
Materials 16 05255 g011
Figure 12. Evolution of microstructure and texture near extrusion die exit before (a) and after (b) sheets forming in TGE process [60].
Figure 12. Evolution of microstructure and texture near extrusion die exit before (a) and after (b) sheets forming in TGE process [60].
Materials 16 05255 g012
Figure 13. (a) Schematic diagrams of ACE, (b) Extrusion physical diagram of Mg alloys; (c) FEM results of Mg alloy processed by ACE [63].
Figure 13. (a) Schematic diagrams of ACE, (b) Extrusion physical diagram of Mg alloys; (c) FEM results of Mg alloy processed by ACE [63].
Materials 16 05255 g013
Figure 14. The microstructure and texture evolutions of the ACE sheets: at (a) 33 mm, (b) 22 mm, (c) 14 mm, (d) 8 mm and (e) 2 mm from die exit; (f) ACE sheet [63].
Figure 14. The microstructure and texture evolutions of the ACE sheets: at (a) 33 mm, (b) 22 mm, (c) 14 mm, (d) 8 mm and (e) 2 mm from die exit; (f) ACE sheet [63].
Materials 16 05255 g014
Table 1. The summary of mechanical properties for Mg alloy sheets processed by different extrusion technologies.
Table 1. The summary of mechanical properties for Mg alloy sheets processed by different extrusion technologies.
Composition
(wt%)
Extrusion
Technologies
SamplesMechanical PropertiesRef.
UTS (MPa)YS (MPa)FE (%)rn
AZ31Conventional extrusion (CE)335.6156.220.02.140.27[67]
45°337.4166.621.02.080.26
90°328.3196.316.42.870.22
AZ31Asymmetric extrusion (ASE)315.4149.516.41.000.34[55]
45°326.4124.723.7
90°344.3135.722.1
AZ31Differential speed extrusion (DSE)352.8179.920.1[58]
45°364.3198.322.8
90°341.5225.018.7
AZ31Normal gradient extrusion(NGE, 45°)342.6151.120.91.960.27[59]
45°345.1152.522.91.870.28
90°349.1182.318.52.430.26
AZ31Transverse gradient extrusion (TGE, 52°)350.1210.322.12.850.26[60]
45°356.9102.130.01.150.53
90°350.7117.226.51.300.45
AZ31Asymmetric porthole die extrusion (APE, 45°)337.6180.821.92.710.22[61]
45°379.5180.826.22.940.29
90°389.9180.825.12.010.34
AZ31/W0Asymmetric material composition extrusion300.9160.318.7[62]
AZ31Asymmetric curve extrusion (ACE)329.5172.219.81.850.26[63]
45°333.4148.624.51.670.29
90°337.6152.621.91.370.30
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

Yang, Q.; Zhang, D.; Peng, P.; Wei, G.; Zhang, J.; Jiang, B.; Pan, F. Asymmetric Extrusion Technology of Mg Alloy: A Review. Materials 2023, 16, 5255. https://doi.org/10.3390/ma16155255

AMA Style

Yang Q, Zhang D, Peng P, Wei G, Zhang J, Jiang B, Pan F. Asymmetric Extrusion Technology of Mg Alloy: A Review. Materials. 2023; 16(15):5255. https://doi.org/10.3390/ma16155255

Chicago/Turabian Style

Yang, Qingshan, Dan Zhang, Peng Peng, Guobing Wei, Jianyue Zhang, Bin Jiang, and Fusheng Pan. 2023. "Asymmetric Extrusion Technology of Mg Alloy: A Review" Materials 16, no. 15: 5255. https://doi.org/10.3390/ma16155255

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